catalytic transformation of ethanol to 1,3-butadiene over mgo/sio2 catalyst

298
Lehigh University Lehigh Preserve eses and Dissertations 2018 Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst William E. Taifan Lehigh University Follow this and additional works at: hps://preserve.lehigh.edu/etd Part of the Chemical Engineering Commons is Dissertation is brought to you for free and open access by Lehigh Preserve. It has been accepted for inclusion in eses and Dissertations by an authorized administrator of Lehigh Preserve. For more information, please contact [email protected]. Recommended Citation Taifan, William E., "Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst" (2018). eses and Dissertations. 4255. hps://preserve.lehigh.edu/etd/4255

Upload: others

Post on 11-Sep-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

Lehigh UniversityLehigh Preserve

Theses and Dissertations

2018

Catalytic Transformation of Ethanol to1,3-Butadiene over MgO/SiO2 CatalystWilliam E. TaifanLehigh University

Follow this and additional works at: https://preserve.lehigh.edu/etd

Part of the Chemical Engineering Commons

This Dissertation is brought to you for free and open access by Lehigh Preserve. It has been accepted for inclusion in Theses and Dissertations by anauthorized administrator of Lehigh Preserve. For more information, please contact [email protected].

Recommended CitationTaifan, William E., "Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst" (2018). Theses and Dissertations.4255.https://preserve.lehigh.edu/etd/4255

Page 2: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

Catalytic Transformation of Ethanol to

1,3-Butadiene over MgO/SiO2 Catalyst

by

William E. Taifan

A Dissertation

Presented to the Graduate and Research Committee

of Lehigh University

in Candidacy for the Degree of

Doctor of Philosophy

in

Chemical Engineering

Lehigh University

May 2018

Page 3: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

ii

Copyright © 2018 by William E. Taifan

Page 4: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

iii

CERTIFICATE OF APPROVAL

Approved and recommended for acceptance as a dissertation in partial fulfillment

of the requirements for the degree of Doctor of Philosophy.

Date

Accepted Date

Dissertation Director:

Date Dr. Jonas Baltrusaitis, Ph.D.

Advisor and Committee Chairperson

Assistant Professor of Chemical

Engineering

Lehigh University

Committee Members:

Date Dr. Israel E. Wachs, Ph.D.

G. Whitney Snyder Professor of

Chemical Engineering

Lehigh University

Date Dr. Hugo Caram, Ph.D.

Professor of Chemical Engineering

Lehigh University

Date Dr. Nicholas C. Strandwitz, Ph.D.

Assistant Professor of Materials

Science and Engineering

Lehigh University

Page 5: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

iv

ACKNOWLEDGEMENTS

This dissertation is made possible by the experiences I have had during my graduate school,

where I have had encounters with wonderful individuals that have been impactful in their

own rights. I am dedicating this page to demonstrate my gratefulness for the

encouragement, stimulation, and motivation that I had from these individuals.

I am eternally grateful for my family’s support for my PhD endeavor. Unsurpassed love,

passion, and attention from both my mother and father have been a driving force to

overcome the potential (mental) barrier at difficult times during my research days. I am

thankful that, if not for my father, I would not have even chosen chemical engineering as

my major. I am thankful that, if not for him, I would not have pursued my doctoral degree.

While my father had been the light that I follow, my mother had been my best supporters

throughout my graduate study. My brother has played an instrumental part in my life, as

we share a lot of secrets that are impossible to share with our parents. These secret-sharing

sessions had been accompanied by advices sharing, where giving him advices had made

me a more mature, better human being. I am grateful of my grandmother and grandfather

for their unwavering support, and for their nagging me to find a wife for myself. Since the

wife is finally here, you rest assured that your first grandson will give you some great

grandchildren. I love you all, and I can never envision completing my PhD without the

support from all of you.

Lehigh has not only given me an opportunity for a graduate study, instead, it has given me

my life partner. The encounter we had in Taylor Gym made my dream of having a spouse

with similar passion came true. If there was a little bright side of living in Lehigh, that

would be you. You are my first in a lot of aspects, such as being a medical emergency

contact during your surgery and taking care of your leg. Those experiences changed me as

a person, for the better. I love you very much, so much that I had decided to commit a

lifetime’s worth of opportunities to experience all the steakhouses and sushi places in Japan

and New York City.

To all of my collaborators. Tomáš Bučko (forgive me if I mistyped the accent on your

name), my first ever collaborator, I can’t imagine how my first ever manuscript would be

possible without your help. I gained a lot of insights on DFT calculation, including few

tricks to stabilize those explosive transition states. Dr. Yuanyuan Li and Prof. Anatoly

Frenkel from Brookhaven National Lab, I’m eternally grateful for giving me a free ride on

your XAS proposal to finally nail down the active sites on my catalyst. The sleepless nights

we spent during Thanksgiving break (yes, during that break), finally paid off and hopefully

the manusript got accepted without a lot of revisions. I’m also acknowledging Dr. Nebojsa

Marinkovic, Nicholas Marcella, Amani Ebrahim, for your help in the beamline.

To my lifting dudes Tony Chang, Henry Choo, Steven Rodriguez, I’m grateful for the

knowledge we share during our lifting sessions. From Candito to Sheiko, from percentages

to RPE. I’m grateful for our meet together, it was a very profound experience. Powerlifting

Page 6: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

v

had been a part of my graduate school life. My love for this sport is a causal relationship

with my research, where the former has been a therapy to the detrimental effects of the

latter. I’m grateful as well that this sport has introduced me to a bunch of wonderful people,

Shane Del Bianco, Johnny Luczkovich, Mark Herndon, and my coach, Samuel Bernstein.

May we all be relentless in our pursuit of strength and power. May our backs arch

beautifully during our bench, may we squat lower than our GPA, and may our deadlifts be

three times our bodyweight.

During the early dark days of the research I have also taken solace in my friendships with

great individuals at this university. With no particular order: George Xu Yan, Benjamin

Moskowitz, Fan Ni, Aaron Zhang, Daniyal Kiyani, Chris Keturakis, Yoona Yang, Leah

Spangler, Sagar Sourav, Chris Curran, Evan Koufos, and Lohit Sharma. My mentors from

Wachs’ lab: Minghui Zhu, Soe Lwin, Jih-Mirn Jeng, and Ivan Santos. Please accept my

apologies for missing your names in this list, as you see I have made efforts to include

everyone’s names, but you all are very meaningful to me.

Lastly I would also like to thank the two Professors that are very influential to me during

my PhD study. Professor Israel E. Wachs, for the meaningful discussion and wonderful

lectures on the operando catalyst characterization, emphasizing the importance of under

reaction condition characterization. I gained a lot of knowledge from our discussions

during the lecture and about my PhD work. My adviser, Jonas Baltrusaitis, with whom I

have formed a long-lasting relationship, though I’m a bit bitter that he couldn’t hood me

on my graduation day. He had given me opportunities to travel and attend the biggest

conferences throughout my PhD study. From him I acquired my knowledge, attitude, and

mindset of a hybrid engineer-scientist. I have grown much since I first came here in terms

of both personal and professional development. I am eternally grateful for your advises and

guidance throughout my graduate study, as they all said, you only have one Ph.D. advisor

for the rest of your life.

Page 7: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

vi

TABLE OF CONTENTS

ACKNOWLEDGEMENTS............................................................................................ iv

TABLES OF CONTENTS ............................................................................................. vi

LIST OF TABLES .......................................................................................................... x

LIST OF FIGURES .......................................................................................................xii

ABSTRACT ..................................................................................................................... 1

CHAPTER 1 | Introduction

1. Background .......................................................................................................... 4

1.1. ETB Reaction Network ............................................................................. 10

1.1.1. Reaction Intermediates and Byproducts ................................................... 11

1.1.2. Proposed Reaction Mechanisms ............................................................... 12

1.2. Catalytic Systems ..................................................................................... 15

1.2.1. Reaction Conditions and Catalytic Performance ...................................... 16

1.2.2. MgO/SiO2 Catalysts ................................................................................. 18

1.2.3. ZrOx –based Catalysts ............................................................................... 24

1.2.4. Other Catalysts ......................................................................................... 26

2. Approach ............................................................................................................ 29

2.1. Approach .................................................................................................. 29

2.2. DFT Calculation ....................................................................................... 30

2.3. In-situ and Operando Spectroscopy ......................................................... 32

2.3.1. Infrared Spectroscopy ............................................................................... 33

2.3.2. UV-Vis Spectroscopy ............................................................................... 34

2.3.3. Operando XANES and EXAFS ................................................................ 35

2.3.4. Temperature-Programmed Reaction Spectroscopy .................................. 36

2.4. HS-LEIS ................................................................................................... 37

2.5. Probe Molecules ....................................................................................... 38

2.6. Product Determination with GC-MS ........................................................ 39 3. Thesis Outline ..................................................................................................... 41

References ....................................................................................................................... 43

CHAPTER 2 | Experimental Methods

1. Introduction ........................................................................................................ 48

2. Computational Details ....................................................................................... 49

2.1. Electronic structure calculations .............................................................. 49

2.2. Structural optimization calculations ......................................................... 49

2.3. Structural model ....................................................................................... 50

2.4. Free-energy calculation ............................................................................ 52

i. Working equations ................................................................................... 52

ii. Partitioning of atomic degrees of freedom in interacting and non-interacting

systems ................................................................................................................. 53

iii. Calculation of harmonic vibrational frequencies ..................................... 54

Page 8: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

vii

3. Experimental Methods ...................................................................................... 55

3.1. Catalyst synthesis ....................................................................................... 55

3.1.1. Synthesis of magnesium oxide, MgO, catalyst ......................................... 55 3.1.2. Synthesis of MgO/SiO2 catalysts .............................................................. 55

3.1.3. Synthesis of promoted wet-kneaded MgO/SiO2 catalysts ......................... 56

3.2. Catalytic reactivity study ........................................................................... 56

3.3. Catalyst characterization ............................................................................ 58

3.3.1. High-sensitivity low energy ion scattering (HS-LEIS) .............................. 58

3.3.2. XRD and BET surface area ........................................................................59

3.3.3. Transition metal concentration measurements ........................................... 59

3.3.4. Scanning transmission electron microscopy .............................................. 60

3.3.5. In-situ spectroscopy ................................................................................... 60

3.3.6. Acid-base characterization using pyridine, NH3, CO2, and methanol as

probe molecules .................................................................................................... 61

3.4. Reaction mechanism study using in-situ DRIFTS spectroscopy and TPRS

............................................................................................................................... 62

3.5. Operando XANES and EXAFS spectroscopy during ethanol reaction to

1,3-BD over Cu- and Zn-promoted MgO/SiO2 catalysts ...................................... 64

References ....................................................................................................................... 65

CHAPTER 3 | Computational Study of Ethanol to 1,3-BD

Reaction Mechanisms

Abstract ........................................................................................................................... 68

1. Introduction ........................................................................................................ 69

2. Computational Results ....................................................................................... 73

2.1. Reaction Pathways .................................................................................. 73

2.1.1. Ethanol dehydrogenation and dehydration ............................................. 73

2.1.2. Aldol condensation ................................................................................. 80

2.1.3. Prins condensation .................................................................................. 85

2.1.4. 1-Ethoxyethanol formation ..................................................................... 88

2.2. Details of the free-energy profiles .......................................................... 89

2.2.1. Elimination/redox reaction of ethanol .................................................... 90

2.2.2. C-C bond formation ................................................................................ 90

2.2.3. Proton transfer ......................................................................................... 91

3. Discussion ............................................................................................................ 92

4. Conclusion ........................................................................................................... 98

Supporting Information .............................................................................................. 100

References ..................................................................................................................... 112

CHAPTER 4 | Combined In-situ DRIFTS and DFT study of

Ethanol to 1,3-BD Reaction Mechanism over MgO/SiO2 catalysts

Abstract .......................................................................................................................... 103

1. Introduction ....................................................................................................... 104

Page 9: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

viii

2. Results and Discussion ...................................................................................... 109

2.1. Catalyst activity and selectivity testing ....................................................... 109

2.2. In-situ DRIFT spectroscopy of MgO based catalyst surface hydroxyl

groups................................................................................................................... 110

2.3. Acid-base characterization of WK (1:1) catalyst using CO2 and pyridine

as probe molecules .............................................................................................. 113

2.4. In-situ DRIFT spectroscopy to monitor hydroxyl group reactivity during

the ethanol, acetaldehyde, crotonaldehyde and crotyl alcohol adsorption and

subsequent reaction on a WK (1:1) catalyst surface ........................................... 115

2.5. In-situ DRIFT spectroscopy of C2 (ethanol, acetaldehyde) and C4

(crotonaldehyde and crotyl alcohol) adsorption and reaction on WK (1:1)

catalyst surface as a function of temperature ...................................................... 119

2.5.1. C2 reactants and intermediates ............................................................... 119

2.5.2. C4 intermediates ..................................................................................... 127

2.6. DFT calculations ethanol, acetaldehyde, crotonaldehyde and crotyl alcohol

vibrational frequencies ........................................................................................ 130

2.7. In-situ DRIFT spectra for the ethanol, acetaldehyde, crotonaldehyde and

crotyl alcohol reaction on a WK (1:1) catalyst surface: the effect of the vapor

phase presence .................................................................................................... 136

3. Conclusions ........................................................................................................ 145

Supporting Information ............................................................................................... 148

References ...................................................................................................................... 151

CHAPTER 5 | Active Sites Determination of MgO/SiO2 Catalysts

for Ethanol to 1,3-BD Reaction

Abstract .......................................................................................................................... 154

1. Introduction ....................................................................................................... 155

2. Results and Discussion ...................................................................................... 160

2.1. Steady state ethanol catalytic conversion to 1,3-BD ................................. 160

2.2. Bulk, surface chemical and structural characterization using XRD, LEIS

and DRIFTS ........................................................................................................ 163

2.3. Temperature-programmed reaction spectroscopy (TPRS) of ethanol on

MgSi-WK ............................................................................................................ 167

2.4. Acid-base characterization using DRIFTS ................................................ 172

2.5. Reactive site persistence during ethanol-to-1,3-BD .................................. 174

2.6. Implications for the structure-activity relationship .................................... 186

3. Conclusions ........................................................................................................ 189

References ...................................................................................................................... 191

CHAPTER 6 | Role of transition metal promoters (Cu, Zn) on

MgO/SiO2 catalyst for Lebedev reaction

Abstract ..........................................................................................................................194

1. Introduction .......................................................................................................195

Page 10: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

ix

2. Computational results ...................................................................................... 199 2.1. Model catalyst selection and analysis ...................................................... 199

2.2. Reactive intermediates ............................................................................. 206

2.3. Potential energy surfaces .......................................................................... 210

3. Experimental results .........................................................................................214 3.1. Catalyst characterization .......................................................................... 214

3.2. Steady state catalytic performance and acid/base chemistry of the

catalyst active sites .............................................................................................. 222

3.3. Active sites under operating conditions ................................................... 226

3.3.1. Temperature programmed infrared spectroscopy measurements (TP-

DRIFTS) ............................................................................................................. 226

3.3.2. In-situ UV-Vis DRS study of MgSi catalysts . 230

3.3.3. Operando XAS studies of Cu, Zn-promoted MgSi catalysts ................... 232

3.3.3.1. Operando XANES and EXAFS of Cu-promoted MgSi catalyst ......... 232

3.3.3.2. Operando XANES and EXAFS of Zn-promoted MgSi catalyst ......... 241

4. Conclusion ......................................................................................................... 245

Supporting Information ............................................................................................... 247

References ...................................................................................................................... 266

CHAPTER 7 | Conclusions and Future Outlook

1. Conclusions......................................................................................................... 270

2. Future Outlook................................................................................................... 273

References ......................................................................................................................275

CURRICULUM VITAE ................................................................................ 276

Page 11: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

x

LIST OF TABLES

Chapter 1

Table 1.1. Catalytic performance of different catalysts studied for one-step ethanol to

1,3-BD conversion. a Contact (residence) time; WHSV (weighted-hourly space

velocity); TOS (time-on-stream); X (ethanol conversion), Y (yield); P (productivity)

......................................................................................................................................... 17

Chapter 3

Table 3.1 Electronic and free energy values of the stationary points calculated at 0 K

and 723 K, respectively .................................................................................................. 75

Table 3.2 Computed forward and reverse reaction barriers and the corresponding

reaction rate constants ..................................................................................................... 95

Chapter 4

Table 4.1. Catalytic activity comparison of WK (1:1) with previously investigated wet-

kneaded synthesized catalysts ......................................................................................... 109

Table 4.2. Surface hydroxyl group vibrational frequencies during ethanol,

acetaldehyde, crotonaldehyde and crotyl alcohol adsorption on WK (1:1) ................... 116

Table 4.3. Vibrational frequencies and their assignments for ethanol, acetaldehyde,

crotonaldehyde and crotyl alcohol adsorption on WK (1:1) ........................................... 121

Table 4.4. Calculated infrared frequencies of ethanol, acetaldehyde, crotonaldehyde

and crotyl alcohol molecules adsorbed on low coordination model MgO surface sites.

Frequencies were calculated using PBE density functional and no scaling to correct for

anharmonicity was applied ............................................................................................. 135

Table S4.1. Calculated infrared frequencies of gas phase ethanol, acetaldehyde, crotyl

alcohol and crotonaldehyde molecules. Frequencies were calculated using PBE density

functional and no scaling to correct for anharmonicity was applied. Experimental

frequencies, except for crotonaldehyde, were obtained from NIST .............................. 149

Chapter 5

Table 5.1 Steady state reactivity of MgO/SiO2 catalysts of different calcination

temperature and preparation method. Reaction was carried out at 450 °C with catalyst

mass of 0.1 g, 55 ml/min total flow rate and pethanol = 2.5 kPa. Selectivity towards major

(by)products ethylene, acetaldehyde and 1,3-BD is reported ......................................... 161

Table 5.2 m/z selection to identify the arising vapor-phase species from TPRS

experiments...................................................................................................................... 168

Page 12: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

xi

Table 5.3. Comparison between observed experimental values of NH3 adsorption on

MgSi-WK catalysts with DFT calculated IR vibrations of NH3 adsorbed on open and

closed acid Mg3C and Mg4C sites. Scaling factor of 0.9854 was applied to the calculated

values and was derived from the gas-phase NH3 experimental and DFT calculated

frequencies....................................................................................................................... 182

Chapter 6

Table 6.1 Different configurations tested for Zn(Cu)/MgO model catalysts. Various

dopant location was chosen between the top and second layer, and compared for energy

and Bader charge. ............................................................................................................203

Table 6.2. Referenced electronic and corrected Gibbs free energy for each species over

MgO, Cu/MgO, and Zn/MgO catalysts. ......................................................................... 208

Table 6.3 Activation energy and thermodynamics consideration for key steps during

ethanol conversion to 1,3-butadiene over MgO, Zn/MgO, and Cu/MgO catalysts......... 211

Table 6.4. Vibrational frequencies in 1600-1400 cm-1 wavenumber range and their

assignments for ethanol, acetaldehyde, crotonaldehyde and crotyl alcohol adsorption

on WK (1:1)11.................................................................................................................. 230

Table 6.5. Best fitting results of Cu catalysts. The structural parameters of standards

were listed for comparison............................................................................................... 233

Table 6.6. Best fitting results for ZnMgSi, ZnMg, ZnO, MgO and Zn. The structural

parameters of standards were listed for comparison. ...................................................... 245

Table S6.1. Peak assignments of surface CO2 species identified on MgSi, CuMgSi, and

ZnMgSi catalysts............................................................................................................. 251

Table S6.2. DFT simulation of NH3 on MgO slab. Simulation was done using VASP,

PBE functionals on 2x2x1 k-point mesh......................................................................... 256

Table S6.3. Redox properties of the MgSi, CuMgSi and ZnMgSi catalysts and

reference MgO obtained from MS measurements. These results have been normalized

to the BET surface area (m2/g) of each catalysts ............................................................ 263

Page 13: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

xii

LIST OF FIGURES

Chapter 1

Figure 1.1. Ethanol production rate increase from 2010 to 2017 (adapted from U.S.

Energy Information and Administration) (left) and ethanol upgrading map to different

highly valued chemicals (right)....................................................................................... 4

Figure 1.2. Proposed reaction mechanisms for ethanol conversion to 1,3-BD: (a)

Toussaint’s aldol condensation; (b) Gruver’s Prins condensation; (c) Cavani’s

carbanion mechanism ......................................................................................................13

Figure 1.3. Operando spectroscopy setup, flow reaction cell temperature/pressure

controller equipped with FTIR, UV-Vis and Raman spectroscopy that enables real-time

online measurement. Output is connected to real-time GC/MS system. Adapted from: http://www.lehigh.edu/operando......................................................................................

. 33

Chapter 2

Figure 2.1 Periodic MgO slab used throughout the calculations. The whited out bottom

layer indicates the atoms whose positions were kept frozen during

calculations...................................................................................................................... 51

Chapter 3

Figure 3.1. Reaction mechanisms proposed for ethanol to 1,3-butadiene; (a)

Toussaint’s generally accepted mechanism, (b) Fripiat’s Prins mechanism, (c)

Ostromislensky’s hemiacetal rearrangement................................................................... 72

Figure 3.2 All stable intermediates and transition states calculated following the

reaction pathways. (1A-1C): ethanol dehydrogenation to acetaldehyde; (2A-2O):

acetaldehyde aldol condensation to 3-hydroxybutanal (acetaldol) followed by proton

transfer to crotonaldehyde; (3A-3G): MPV (Meerwein–Ponndorf–Verley) reduction of

crotonaldehyde to 1,3-butadiene; (4A-4K): acetaldol MPV reduction to butadiene; (5A-

5C): ethanol dehydration to ethylene; (6A-6E iii 3): Prins condensation of acetaldehyde

and ethylene; (7A-7E): ethanol and acetaldehyde nucleophilic addition reaction

(Ostromislensky’s hemiacetal rearrangement)................................................................ 79

Figure 3.3 Free-energy profiles for (a) ethanol dehydrogenation to form acetaldehyde

and (b) ethanol dehydration to ethylene...........................................................................79

Figure 3.4 Free-energy profiles for aldol condensation pathway................................... 81

Figure 3.5 Free-energy profiles for the MPV reduction of the resulting molecule from

aldol condensation. Red pathway indicates subsequent proton transfer of acetaldol

followed by MPV reduction of the crotonaldehyde; Blue pathway shows the direct

MPV reduction of the resulting acetaldol....................................................................... 84

Page 14: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

xiii

Figure 3.6 Free-energy profiles for the Prins condensation between ethylene and

acetaldehyde. Red pathway indicates a typical route of Prins condensation; Blue

pathway shows an additional proton diffusion step in between the reaction steps; Black

pathway shows the unlikely formation of MEK............................................................. 87

Figure 3.7 Free-energy profile for ethanol and acetaldehyde nucleophilic addition

reaction............................................................................................................................ 89

Chapter 4

Figure 4.1. Main reaction mechanism proposed for ethanol to 1,3-butadiene via

Toussaint’s aldol condensation ....................................................................................... 105

Figure 4.2. Conversion (●) and selectivity of main products (■ acetaldehyde; ▲

ethylene; ♦ 1,3-butadiene) at different WHSV. Reaction conditions: T=723 K, Qtot =

50 cm3/min, Mcat=0.2 g, P0EtOH = 2.72; 3.77; 5.15; 6.96 kPa........................................... 110

Figure 4.3. In-situ DRIFTS spectra acquired of dehydrated (temperature programmed

to 773 K at 10 oC/min under air and cooled down to 100 oC) MgO, MgO WK (1:1)

catalysts and SiO2. Only hydroxyl region of 3800 to 3200 cm-1 is shown. Spectra are

acquired at 100 °C............................................................................................................ 111

Figure 4.4. DRIFTS spectra of adsorbed (a) CO2 and (b) pyridine on WK (1:1) catalyst

at different temperatures to probe the catalyst’s basicity and acidity at relevant

temperatures..................................................................................................................... 114

Figure 4.5. In-situ DRIFTS spectra in the hydroxyl group region of 3800 – 3200 cm-1

acquired of ethanol, acetaldehyde, crotonaldehyde and crotyl alcohol on WK (1:1)

catalyst. Sample vapor was adsorbed on the sample surface and temperature ramped

up from 373 to 723 K while spectra being recorded. In-situ DRIFTS dehydrated

catalyst spectrum at 100 °C was used as a reference....................................................... 118

Figure 4.6. In-situ DRIFTS spectra acquired of ethanol on WK (1:1) catalyst in 3200

to 1000 cm-1 spectral region. Ethanol was adsorbed on the sample surface and

temperature increased from 373 to 723 K while spectra being recorded. In-situ

DRIFTS spectra of the sample surface with no adsorbate present at every corresponding

temperature were used for reference. In-situ DRIFTS spectra acquired for ethanol

adsorbed on MgO are shown in the inset for 1200 to 1000 cm-1 spectral region............ 123

Figure 4.7. In-situ DRIFTS spectra acquired of acetaldehyde on WK (1:1) catalyst in

3200 to 1000 cm-1 spectral region. Acetaldehyde was adsorbed on the sample surface

and temperature increased from 373 to 723 K while spectra being recorded. In-situ

DRIFTS spectra of the sample surface with no adsorbate present at every corresponding

temperature were used for reference. In-situ DRIFTS spectra acquired for acetaldehyde

adsorbed on MgO are shown in the inset for 1200 to 1000 cm-1 spectral

region............................................................................................................................... 126

Page 15: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

xiv

Figure 4.8. In-situ DRIFTS spectra acquired of crotonaldehyde on WK (1:1) catalyst

in 3200 to 1000 cm-1 spectral region. Crotonaldehyde was adsorbed on the sample

surface and temperature increased from 373 to 723 K while spectra being recorded. In-

situ DRIFTS spectra of the sample surface with no adsorbate present at every

corresponding temperature were used for reference........................................................ 127

Figure 4.9. In-situ DRIFTS spectra acquired of crotyl alcohol on WK (1:1) catalyst in

3200 to 1000 cm-1 spectral region. Crotyl alcohol was adsorbed on the sample surface

and temperature increased from 376 to 723 K while spectra being recorded. In-situ

DRIFTS spectra of the sample surface with no adsorbate present at every corresponding

temperature were used for reference................................................................................ 128

Figure 4.10. PBE optimized structures of ethanol (I), acetaldehyde (II), its enolate

conformation (II), crotonaldehyde (IV), crotyl alcohol (V) and 1,3-butadiene (VI) on

MgO surface low coordination Mg3cO4c or Mg3cO5c surface sites. Numbers refer to the

particular steps in catalytic transformation cycle shown in Figure 4.1........................... 134

Figure 4.11. In-situ DRIFTS spectra acquired of ethanol on WK (1:1) catalyst. Ethanol

was adsorbed on the sample surface, flown continuously and temperature increased

from 376 to 723 K while spectra being recorded. In-situ DRIFTS spectrum of the

sample surface with adsorbed ethanol present at 373 K was used for reference............. 139

Figure 4.12. In-situ DRIFTS spectra acquired of acetaldehyde on WK (1:1) catalyst.

Acetaldehyde was adsorbed on the sample surface, flown continuously and temperature

increased from 376 to 723 K while spectra being recorded. In-situ DRIFTS spectrum

of the sample surface with adsorbed acetaldehyde present at 373 K was used for

reference........................................................................................................................... 140

Figure 4.13. In-situ DRIFTS spectra acquired of crotonaldehyde on WK (1:1) catalyst

under ethanol vapor flow. Crotonaldehyde was adsorbed on the sample surface,

flushed with inert gas and ethanol was introduced under continuous flow with

temperature increased from 376 to 723 K while spectra being recorded. In-situ

DRIFTS spectrum of the sample surface with adsorbed crotonaldehyde at 373 K was

used for reference. For comparison, 523 K spectrum of crotonaldehyde adsorbed with

no gas phase present is shown in red dotted line............................................................. 143

Figure 4.14. Complete surface reaction scheme on ethanol reaction over MgO/SiO2

catalyst. (I) Crotonaldehyde, (II) adsorbed crotyl alcohol, (III) 1,3-butadiene, (IV) 2,4-

hexadienal, (V) paraldehyde, (VI) metaldehyde. ............................................................ 147

Figure S4.1. In-situ spectroscopy of ethanol on MgO catalyst. Ethanol was adsorbed

on the sample surface and temperature ramped up from 373 to 723 K while spectra

being recorded. Subtracted spectra are shown. Spectra are offset for clarity................ 148

Figure S4.2. In-situ spectroscopy of acetaldehyde on MgO catalyst. Acetaldehyde was

adsorbed on the sample surface and temperature ramped up from 373 to 723 K while

spectra are being recorded. Spectra are offset for clarity................................................ 148

Page 16: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

xv

Chapter 5

Figure 5.1. Possible combination of metal atoms that act as Lewis acid sites: A: Mg3C

(open), B: Mg3C (closed), C: Mg4C (closed), D: Mg4C (open), E: Mg5C (open), F: Mg5C

(closed)............................................................................................................................. 159

Figure 5.2. Catalytic activity of MgSi-WK between 350-450°C. Inset: Arrhenius plot

of ethylene and 1,3-BD. Catalyst mass = 0.1 gr, total flow rate = 55 ml/min, pethanol =

2.5 kPa. ....................................................................................... ................................... 163

Figure 5.3. Comparison of XRD patterns of MgSi-WK and MgSi-IWI. WK with

different oxide ratios are overlaid for comparison. ......................................................... 164

Figure 5.4. Depth-profile of (a) MgSi-IWI and (b) MgSi-WK as probed using HS-

LEIS. HS-LEIS spectra of layer by layer sputtering of catalyst surface are shown in the

inset.................................................................................................................................. 165

Figure 5.5. Left: Comparison of OH groups of MgSi-WK and MgSi-IWI as probed by

in-situ dehydrated DRIFTS experiments; right: OH groups of WK catalysts with

different oxide ratios........................................................................................................ 167

Figure 5.6. TPRS spectra of ETB reaction over MgSi-WK with ethanol as the feed

(left) and acetaldehyde as the feed (right). EtOH: ethanol; AA: acetaldehyde; CA:

crotonaldehyde; C-OH: crotyl alcohol............................................................................. 169

Figure 5.7. TPRS spectra of ETB reaction over MgSi-WK with ethanol and

acetaldehyde as the coreactants. Acetaldehyde is pre-adsorbed on the surface and

temperature ramp is under ethanol flow.......................................................................... 171

Figure 5.8. Acid-base characterization of MgSi-IWI and MgSi-WK catalysts probed

using CO2 (left) and pyridine (right). Spectra at high temperature (450°C) and low

temperature (100°C) are shown....................................................................................... 173

Figure 5.9. In-situ acid-base characterization of MgSi-WK catalyst before and after

ethanol adsorption at 100°C and reaction at 200°C using CO2 (left) and pyridine

(right)............................................................................................................................... 174

Figure 5.10. Acid-base poisoning reactivity testing using (a) CO2, (b) propionic acid,

and (c) NH3 to determine the role of each site during ethanol conversion to 1,3-BD over

WK-800 MgO/SiO2 catalyst. Reactions are carried out at 425 °C, mcat = 0.1 g, pethanol =

2.5 kPa, total flow = 55 ml/min. All formation rates are normalized to initial 1,3-BD

formation rate................................................................................................................... 178

Figure 5.11. Productivity of (a) 1,3-BD, (b) ethylene, and (c) acetaldehyde of Na-

poisoned MgSi-WK catalysts between 350-450°C. Catalyst mass = 0.1 gr, total flow

rate = 55 ml/min, pethanol = 2.5 kPa................................................................................... 181

Page 17: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

xvi

Figure 5.12. Bottom: DRIFTS characterization of Na-doped MgSi-WK using (a) CO2

and (b) NH3. Spectra are taken at 100°C after extensive evacuation with N2. Top: (a)

CO2 desorption spectra of 1000 ppm Na-doped MgSi-WK at 100, 300, and 450°C and

(b) NH3 desorption on 0 ppm and 250 ppm Na-doped MgSi-WK at 300°C. Spectral

subtraction was done using the spectra of the dehydrated catalysts at respective

temperatures as the background....................................................................................... 184

Figure 5.13. (a) MgO periodic model used for DFT simulation of NH3 adsorption on

MgO Lewis acid sites: (b) Mg3C, closed, (c) Mg4C, closed, (d) Mg3C, open, (e) Mg4C,

open. Multiple possible adsorption sites, i.e. kink (Mg3CO4C), edge (Mg4CO4C), and

planar (Mg5CO5C) are highlighted.................................................................................... 185

Figure 5.14. Schematic diagram to show the presence of various sites investigated with

NH3 and CO2 DRIFTS experiments. The basic sites (orange) are shown in the figure as

both Brønsted base (OH) and Lewis site (electron accepting oxygen atoms), and acid

sites (blue) are represented as Brønsted acid sites (H) and Lewis acid sites (electron

donating magnesium and silicon atoms).......................................................................... 186

Figure 5.15. Representation of the role of basic sites during ethanol conversion to

acetaldehyde. Top figure represents dehydrated (pretreated) catalyst; bottom figure

demonstrates the absence of bicarbonate when CO2 is adsorbed in-situ after reaction at

200 °C.............................................................................................................................. 188

Chapter 6

Figure 6.1. Local structure analysis of (a)MgO, (b)Cu-MgO, and (c)Zn-MgO. The

Bader atomic charge on each atom is indicated by the boldfaced numbers.................... 204

Figure 6.2. All stable intermediates and transition states calculated following the

reaction pathways. (1A-1C): ethanol dehydrogenation to acetaldehyde; (2A-2C):

ethanol dehydration to ethylene; (3A-3C): C-C bond formation step in acetaldehyde

aldol condensation to 3-hydroxybutanal (acetaldol); (4A-4C): C-C bond formation step

in Prins condensation of acetaldehyde and ethylene. Calculations are carried over

Zn/MgO model catalysts (prefix: Zn), and Cu/MgO model catalysts (prefix: Cu)......... 207

Figure 6.3. Potential energy surface for ethanol (a)dehydrogenation and

(b)dehydration over MgO, Zn/MgO, and Cu/MgO catalysts. (●)MgO, (■) Cu-MgO, (♦)

Zn-MgO........................................................................................................................... 212

Figure 6.4. Potential energy surface for first C-C bond formation via (a) acetaldehyde

aldol condensation and (b) Prins reaction between acetaldehyde and ethylene over

MgO, Zn/MgO, and Cu/MgO catalysts........................................................................... 213

Figure 6.5. Comparison of XRD patterns between CuMgSi, ZnMgSi, and MgSi......... 215

Figure 6.6. In-situ dehydrated DRIFTS of OH region of MgSi, CuMgSi, and ZnMgSi.

Spectra were taken at 100°C under He flow after pretreatment at 500°C for 1 hour.

Spectra were offset for clarity.......................................................................................... 216

Page 18: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

xvii

Figure 6.7. In-situ UV-Vis DRS spectra of (a) dehydrated CuMgSi catalyst referenced

with Cu/MgO (CuMg), Cu/SiO2 (CuSi), CuO, and MgSi; (b) dehydrated ZnMgSi

catalyst referenced with Zn/MgO (ZnMg), Zn/SiO2 (ZnSi), ZnO, and MgSi. Inset: UV-

Vis spectra of different loadings of Zn on MgO/SiO2

catalysts........................................................................................................................... 217

Figure 6.8. Scanning Transmission Electron Microscopy images of ZnMg, ZnMgSi,

CuMg and CuMgSi samples. Energy Dispersive Spectroscopy profiles (smoothed) are

also provided. Small ZnO nanoparticles are shown in ZnMgSi with red

arrows.............................................................................................................................. 220

Figure 6.9. Productivity comparison of 1,3-BD (■), ethylene (●), and acetaldehyde

(▲) over (a) MgSi, (b) CuMgSi, and (c) ZnMgSi. Dotted lines are meant to guide the

eyes. Insets: Arrhenius plots to show apparent activation energies of the three

(by)products. Reactions are carried out between 325 - 450°C, mcat = 0.1 g, pethanol = 1.8

kPa, total flow = 55 ml/min............................................................................................. 224

Figure 6.10 Evolution of each peak during in-situ temperature-programmed ethanol

DRIFTS over (a) MgSi, (b) CuMgSi, (c) ZnMgSi. Insets: original spectra of ethanol

DRIFTS from where the peaks were deconvoluted......................................................... 227

Figure 6.11. In-situ UV-Vis DRS under constant ethanol flow over (a) CuMgSi and

(b) ZnMgSi...................................................................................................................... 231

Figure 6.12. Normalized XANES spectra of CuMg, CuSi, and CuMgSi (a) and Cu foil,

CuO, Cu2O, and CuMg (b). Inset: Cu K-edge k2-weighted EXAFS data of

corresponding spectra. XANES spectra in Figure 6.12(a) are offset vertically for

clarity............................................................................................................................... 232

Figure 6.13. Normalized temperature-programmed operando XANES spectra of

CuMgSi catalyst under He flow (top) and ethanol flow (bottom). Inset: enlarged region

of the pre-edge features to elucidate changes at different temperature........................... 235

Figure 6.14. Normalized time-resolved operando XANES spectra of CuMgSi catalyst

under ethanol flow at 400°C. Inset: enlarged region of the pre-edge features to elucidate

changes at different temperature...................................................................................... 236

Figure 6.15. Coordination number changes during reaction of ethanol to 1,3-BD over

CuMgSi............................................................................................................................ 237

Figure 6.16. XANES spectra of the simulated CuO Model 1: Cu in a local evironment

surrounded by 6 oxygen atoms and Model 2: Cu in a local environment surrounded by

4 oxygen atoms................................................................................................................ 239

Figure 6.17. (a) Normalized XANES spectra of ZnMg, ZnSi, ZnMgSi, Zn foil, and

ZnO. Inset: Zn K-edge k2-weighted EXAFS data of corresponding spectra. (b) Fourier

transforms of the EXAFS spectra of ZnMg, ZnO, and Zn

foil.................................................................................................................................... 242

Page 19: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

xviii

Figure 6.18. Normalized temperature-programmed operando XANES spectra of

ZnMgSi catalyst under He flow (a) and ethanol flow (b). Inset: enlarged region of the

pre-edge features to elucidate changes at different temperature. (c) Temperature-

induced change in coordination number of Zn-Mg and Zn-O bonds during the

reaction............................................................................................................................ 243

Figure S6.1. XRD patterns of (a) Zn/MgO and (b) Zn/SiO2 at different

loadings............................................................................................................................ 247

Figure S6.2. XRD patterns of (a) Cu/MgO and (b) Cu/SiO2 at different

loadings............................................................................................................................ 247

Figure S6.3. In-situ DRIFTS of OH region of dehydrated MgSi catalysts references at

100°C for Cu-promoted (left) and Zn-promoted (right). Spectra are offset for

clarity............................................................................................................................... 248

Figure S6.4. Tauc plot of CuO (left) and deconvoluted Cu species of CuMgSi catalyst

(right) to determine the edge energy/band gap (E0) for correlation with number of Cu

coordination..................................................................................................................... 248

Figure S6.5. In-situ UV-Vis difference spectra of oxidative dehydration of (a) CuMgSi

and (b) ZnMgSi................................................................................................................ 248

Figure S6.6. Poisoning reactivity testing using CO2 to determine the role of basic sites

during ethanol conversion to 1,3-BD over (a) MgSi, (b) CuMgSi, and (c) ZnMgSi.

Reactions are carried out at 400 °C, mcat = 0.1 g, pethanol = 2.5 kPa, total flow = 55

ml/min. All formation rates are normalized to initial 1,3-BD formation

rate................................................................................................................................... 249

Figure S6.7. CO2 Temperature Programmed-DRIFTS on (a) MgSi, (b) CuMgSi, and

(c) ZnMgSi....................................................................................................................... 253

Figure S6.8. Poisoning reactivity testing using propionic acid to determine the role of

basic sites during ethanol conversion to 1,3-BD over (a) MgSi, (b) CuMgSi, and (c)

ZnMgSi. Reactions are carried out at 400 °C, mcat = 0.1 g, pethanol = 2.5 kPa, total flow

= 55 ml/min. All formation rates are normalized to initial 1,3-BD formation

rate....................................................................................................................................254

Figure S6.9. Poisoning reactivity testing using NH3 to determine the role of acid sites

during ethanol conversion to 1,3-BD over (a) MgSi, (b) CuMgSi, and (c) ZnMgSi.

Reactions are carried out at 400 °C, mcat = 0.1 g, pethanol = 2.5 kPa, total flow = 30 ml/min

(without NH3), 55 ml/min (with NH3). All formation rates are normalized to initial 1,3-

BD formation rate. NH3 desorption spectra on MgSi catalysts at 100°C are shown in

(d) .................................................................................................................................... 255

Page 20: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

xix

Figure S6.10. DRIFTS spectra in the C-H stretching (left) and bending (right) region

of methanol desorption under He flow on unpromoted (top) and promoted (bottom)

catalysts............................................................................................................................ 257

Figure S6.11. Online MS analysis during operando methanol DRIFTS of CuMgSi,

ZnMgSi, MgSi and reference MgO................................................................................. 260

Figure S6.12. In-situ UV-Vis DRS of ethanol reaction on undoped MgO/SiO2 catalyst.

Difference spectra is shown, where catalyst spectra at 100°C with chemisorbed ethanol

is used as a reference........................................................................................................264

Figure S6.13. R-space EXAFS spectra of CuMg catalyst, in comparison to Cu foil,

CuO, and Cu2O................................................................................................................ 264

Figure S6.14. Corresponding MS data of in-situ XANES-EXAFS for ethanol to 1,3-

BD over (a) CuMgSi, (b) ZnMgSi................................................................................... 265

Page 21: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

1

Abstract

Increasing concerns regarding global warming, which is caused by growing CO2

emissions, have led to efforts focused on discovering alternatives to petroleum for energy

and commodity chemical production. (Bio)ethanol has been seen as a platform molecule

with increasing production and versatility for upgrading to various high-value fuels and

chemicals. Among those high-value chemicals is 1,3-butadiene (1,3-BD), which has

demonstrated widespread applications in polymer synthesis and as an organic chemistry

intermediate. Its conventional methods of production rely on oil as a feedstock, hence

suggesting the need for alternative and more sustainable routes. Interest in the catalytic

conversion of ethanol to 1,3-BD, introduced in the 1940s by Lebedev, has been revived

and is now focused on the development of selective catalysts, thus minimizing the need for

the high cost separation between 1,3-BD and other (by)products, such as C2 and C4 olefins

and oxygenates. The main components of the catalyst for this system are MgO and SiO2,

where its reactivity and selectivity depend heavily on the method of preparation. This

system is still at an early stage of development, with a lot of disagreements on structure of

the catalyst, optimum ratio of Mg:Si, reactive intermediates, reaction mechanisms, and

kinetics.

Reaction mechanism was studied intensively using both theoretical (DFT) and

experimental (spectroscopy) methods. Initial screening of the reaction mechanism using

DFT with MgO defect site, i.e. kink, demonstrated that aldol condensation is more viable

thermodynamically than Prins condensation. In the reaction mechanism, dehydrogenation

of ethanol to acetaldehyde, an important reactive intermediate, is shown to be the rate-

determining step (RDS) of the reaction. Comparison of the potential energy barrier also

Page 22: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

2

shows that acetaldol, the product of acetaldehyde self-aldolisation, dehydration competes

with its hydrogenation with an ethanol molecule. This mechanistic study is also supported

by comprehensive in-situ DRIFTS. MgO/SiO2 catalyst is synthesized using a wet-kneading

method, with equivalent oxide mass ratio and thoroughly characterized with HS-LEIS,

DRIFTS, and XRD. Chemical probing was also done with different probe molecules, such

as pyridine, NH3, CO2, and methanol. Combination of several reactants and intermediate

shows that acetaldehyde is spontaneously transformed to crotonaldehyde under constant

reactant flow, while in-situ ethanol DRIFTS requires contribution from the gas-phase

ethanol to make 1,3-BD. Furthermore, the crotonaldehyde does not transform to 1,3-BD

under inert flow, it requires the presence of ethanol to complete the transformation to 1,3-

BD.

The resulting catalyst was extensively probed and characterized, revealing a silica-

rich surface, where comparison with incipient wetness impregnation catalyst shows a rather

Mg-rich surface. Surface silicate that is formed is confirmed by in-situ DRIFTS, where

new OH groups were formed. The basicity of the catalyst also varies significantly with

different methods of preparation and calcination temperature. All strong, medium, and

weak basic sites were found on the catalysts surface. More superior performance, however,

is shown to be enforced by lower amount of strong basic sites. Ammonia probing reveals

the presence of both open and closed Lewis acid sites (LAS) and limited amount of

Brønsted acid sites (BAS). Pyridine, on the other hand, could not identify any BAS, which

is due to its larger molecule size. This further demonstrates that the LAS on the catalyst is

much more accessible than the BAS.

Page 23: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

3

Promotion of the catalyst with transition metal was shown to have a significant

enhancement on the reactivity. Since the RDS was determined to be the dehydrogenation

of ethanol, transition metal sites lower this barrier, and shift the RDS. Zn and Cu, two very

promising ethanol dehydrogenation catalysts, were separately impregnated on the

uncalcined wet-kneaded MgO/SiO2 support at low loadings, 2.5 and 1%, respectively. The

catalysts were thoroughly characterized using in-situ UV-Vis, methanol operando

DRIFTS, in-situ XANES and EXAFS, TEM, TPSR, and in-situ DRIFTS. Cu(II) exists as

a surface species coordinated in a tetrahedral geometry, where it has 0.8 (or ~1) nearest

neighbor, i.e. number of Cu-O-Cu bonds. The transition metal also possesses Cu-O-Mg

bond, hinting to formation of solid solution. Similar interaction was also observed for Zn,

suggesting the stronger interaction with Mg, instead of Si. This structural change affects

the basicity and acidity of the catalyst. Both CO2 and methanol probing with DRIFTS show

that the promoted catalysts have less affinity with CO2, while the BAS was eliminated,

replaced with another distinct LAS. Redox capability was also modified, shown by the

enhanced strength of the redox site in expense of its reduced quantity. During the reaction,

Cu(II) is reduced to Cu(0) via an intermediate Cu species, before the catalyst deactivates

after long hours of experiment. Zn, on the other hand, maintained its structure even after

extensively tested.

Page 24: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

4

Chapter 1

Introduction

1. Background .......................................................................................................... 4

1.1. ETB Reaction Network ............................................................................. 11

1.1.1. Reaction Intermediates and Byproducts ................................................... 11

1.1.2. Proposed Reaction Mechanisms ............................................................... 12

1.2. Catalytic Systems ..................................................................................... 15

1.2.1. Reaction Conditions and Catalytic Performance ...................................... 16

1.2.2. MgO/SiO2 Catalysts ................................................................................. 18

1.2.3. ZrOx –based Catalysts ............................................................................... 24

1.2.4. Other Catalysts ......................................................................................... 26

2. Approach ............................................................................................................ 29

2.1. Approach .................................................................................................. 29

2.2. DFT Calculation ....................................................................................... 31

2.3. In-situ and Operando Spectroscopy ......................................................... 32

2.3.1. Infrared Spectroscopy ............................................................................... 33

2.3.2. UV-Vis Spectroscopy ............................................................................... 34

2.3.3. Operando XANES and EXAFS ................................................................ 35

2.3.4. Temperature-Programmed Reaction Spectroscopy .................................. 36

2.4. HS-LEIS ................................................................................................... 37

2.5. Probe Molecules ....................................................................................... 38

2.6. Product Determination with GC-MS ........................................................ 39 3. Thesis Outline ..................................................................................................... 41

References ....................................................................................................................... 43

1. Background

The growing environmental concerns caused by the increasing CO2 emissions have

incentivized endeavors on discovering alternatives for energy and chemical production.

While potential alternatives such as wind, solar, and nuclear had been able to partially

replace the need for power generation, biomass remains one the only options to mitigate

the petroleum consumption in chemical production section.1 Biomass valorization had

Page 25: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

5

been extensively studied, with focuses on lignin depolymerization, sugar modification, and

hemicellulose fermentation to ethanol.2,3 The target molecules varied from aromatics,

alcohols, to new molecules that were envisioned to replace the incumbent from petroleum.2

While the fight for biomass upgrading is still far from over, (bio)ethanol has presented a

very interesting alternative, due to its abundance3 and its versatility to be upgraded to

different other higher-valued platform molecules or commodity chemicals. Figure 1.1

shows the ever-increasing production of ethanol and its upgradeability to different

molecules.

Ethanol upgrading to higher-valued chemicals has recently been pursued.4,5 It is

fairly reactive, and the fact that it has two carbons makes it relatively selective to even-

numbered carbon containing molecules. Different target were investigated, such as

hydrogen,5–7 n-butanol,8–13 ethylene and diethyl ether (DEE),14–20 propylene,21–25

isobutene,26–28 ethylene oxide,29,30 acetaldehyde,31–35 ethyl acetate,36–42 and 1,3-butadiene

(1,3-BD). Steam reforming had also been explored to replace the current energy-intensive

methane steam reforming processes. The ethanol to chemical processes is different from

Figure 1.1. Ethanol production rate increase from 2010 to 2017 (adapted from U.S.

Energy Information and Administration) (left) and ethanol upgrading map to different

highly valued chemicals (right).

Page 26: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

6

methanol to olefin, where the latter is initiated by carbon pooling mechanism, activating

the surface methoxide and breaking the stable C-H bond. Most of the proposed mechanisms

for ethanol upgrading involve acetaldehyde formation, which can be activated by enolate

formation. This reaction itself opens a new pathway for the C-C bond formation, which is

also extensively studied.43

Hydrogen is a very clean fuel energy source, where its application will give off

water as the only product.44 Current hydrogen production is dominated by methane steam

reforming over Ni catalyst, which is also followed by water-gas shift reaction.45 However,

such reaction generally applies severe condition, 700-1100 °C, leading to the very high

capital costs for its facilities.45 Another pathway to make hydrogen is the photocatalytic

process, where water is split into hydrogen and oxygen. However, charge recombination

and thermodynamics stand as a major obstacle in achieving respectable yield.46 Ethanol

steam reforming, on the other hand, possesses a major advantage in the much lower

reaction temperature, <600 °C, some of the processes even had reaction temperatures of

250 °C.5–7 Partial oxidation of ethanol, i.e. autothermal reforming, is also another

alternative that is being investigated, due to it being much less energy intensive than the

endothermic steam reforming. Major challenge in this reaction remains the expected

carbonaceous deposit on the catalyst surface, which will lead to catalyst deactivation. This

carbonaceous deposit, however, can be mitigated to a certain extent by using suitable

supports, such as MgO, ZnO, CeO2, and La2O3.6

Other endeavors had been focused on valorizing ethanol into ethylene, propylene,

ethylene oxide, 1,3-BD, and n-butanol, which are among the top 30 industrial organic

chemicals based on weight produced in the USA.1 Ethylene is typically produced from oil

Page 27: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

7

cracking, which also produces 1,3-BD as the byproduct. Popular alternative of this process

has been the dehydrogenation of ethane, due to the increasing production of ethane from

both shale gas and the byproduct of the paraffin from the oil cracking.4 Ethanol dehydration

was touted as a possible route for the production of bioethylene, where it has recently been

made economically possible due to the recent advances in the heterogenous catalysts, as

well as the lowered ethanol price.4 The dehydration of ethanol is carried out over solid acid

catalysts, such as SAPO-34,17 γ-Al2O3,14–17 and zeolites.14,17,18 γ-Al2O3 is the most stable

catalyst, whereas other catalysts were found to give higher ethylene yield and operate at

lower temperature. Promotion with transition metals, such as Ni17 and Mo,18 were found to

have prolonged the catalysts’ life, where the latter was shown to be more of a sacrificial

transition metal oxide to be reduced during the reaction.18 Lewis acid sites (LAS) were

observed to be the main site for ethanol conversion, while ethylene production was

maximized by the presence of medium and weak acid sites.16,17 Water, byproduct of the

ethanol dehydration, however, plays an important role in deactivating the catalyst, since it

was found to block the neighboring site, which prevents the C-H bond breaking to make

the C=C bond.19 The major challenge remains in limiting the bimolecular dehydration route

to make diethyl ether, which is much more thermodynamically favored than ethanol. This

can be done by using catalyst with confinement effects.20

Another C2 molecule that can be directly synthesized from ethanol is acetaldehyde.

This molecule had traditionally been used as an intermediate that is further converted to

other chemicals, mainly to acetic acid. However, this process was largely abandoned due

to the more selective Monsato and Cativa processes from methanol.47 Production of

acetaldehyde from ethanol follows two routes, partial oxidation and non-oxidative

Page 28: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

8

dehydrogenation. The latter is becoming more attractive due to the production of hydrogen

as its byproduct, even though the presence of oxygen significantly enhances the activity.

Non-oxidative dehydrogenation of the catalyst was studied with different Cu catalysts,

including Cu/SiO2,31 Cu-Cr2O4,

32,33 and CuO/RHA (rice husk ash).34,35 The presence of

chromium stabilized the catalyst by preventing sintering during both reduction and

reaction.32,33 Significant improvement of acetaldehyde yield was achieved when Cu was

supported on rice husk ash, a silica-rich byproduct of domestic agriculture, resulting in

lowered Cu particle size and high surface area catalyst.34,35

Related to acetaldehyde, ethyl acetate is another product of ethanol upgrading that

involves the acetaldehyde as the reactive intermediate. Important catalysts that were

developed are Cu-Zn-Zr-Al-O, 36–38 supported Pd catalysts,39 Au/TiO2,40 and supported Cu

catalysts.41 The presence of different metals were reported to have different effects on the

catalyst.42 Zr and Al, for instance, enhanced the conversion of ethanol, with Zr favored

ethyl acetate, while Al favored dehydration products such as DEE and methyl ethyl ketone

(MEK). Collaborative effects of Zn and Zr were touted to limit the MEK, which is an

unwanted byproduct, while Cu increased the dehydrogenation reaction, resulting in high

ethyl acetate yield.42 There is still a lot of room for improvement for this system, since the

reaction mechanism is yet to be proven. Rational design of the catalysts is still far from

reach, shown by the previous investigators’ attempts to use various different transition

metals in their catalysts.

C4 molecules that are upgraded by creating new C-C bond, such as isobutene, n-

butanol, and 1,3-BD, are very attractive due to its higher value and its multiple applications.

Isobutene is an important commodity chemical, in particular as an additive to jet fuel and

Page 29: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

9

as a raw material for various polymers. Pathway from ethanol was recognized by

converting it into acetone, and further converting acetone into isobutene. The highlighted

two-step reaction, however, requires basic and acidic catalysts. Bifunctional catalysts

containing balanced amount of base/acid sites were synthesized at Pacific Northwest

National Laboratory (PNNL).26–28 The catalysts, possessing two different sites, were able

to catalyze the two reactions in one pot, with basic and redox sites catalyzing the ethanol

to acetone, while the acid sites condensing the acetone into isobutene. Detailed study of

the catalysts revealed the importance and detrimental effects of Brønsted acid sites (BAS),

which catalyzed the second reaction, as well as isomerizing and polymerizing isobutene

into other butenes and coke.28

N-butanol has a lot of advantages over ethanol as a drop-in fuel. It has higher

solubility in gasoline (longer chain) and significantly higher calorific value (29.2 vs 19.6

MJ/dm3). Upgrading ethanol to n-butanol had been a focus of several research groups,8–13

and had converged into two classes of catalysts: hydroxyapatite and hydrotalcite. N-

Butanol has been traditionally produced from aldol condensation of acetaldehyde followed

by the catalytic hydrogenation and from oxo process.47 Revisiting the process is of

paramount interest due to the lowered ethanol price. Just like isobutene, the catalyst needs

to be bifunctional in nature. Very early attempts used combination of strong basic and

acids, such as alkali cation-exchanged zeolites48 and Na-promoted zirconia.49 Magnesia-

based catalysts present a unique class of bifunctional catalysts, in which they do not

necessarily possess strong basic sites, but still provide weak acid sites. Magnesia alone was

observed to give respectable n-butanol yield,8,10 while combination with alumina gave a

hydrotalcite structure, which increased the activity significantly.50–52 More improvement

Page 30: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

10

was achieved by hydroxyapatite catalysts, calcium phosphate material, which was done by

tuning the Ca/P ratio, as stronger basic sites were shown to increase the n-butanol yield.53

Latest works had focused on the study of the material, where kinetic models and active

sites were determined using combination of several methods, such as SSITKA and in-situ

titration with probe molecules.12,13

Among all of the higher-valued target molecules, 1,3-BD represents the most

interesting target molecule. Back in the 1940s, both the US and the USSR produced 1,3-

BD from ethanol via two-step54,55 and one-step processes,56 respectively. In the two-step

process, ethanol was first converted into acetaldehyde before the product and a separate

ethanol feed were flowed to a second reactor containing a supported tantalum oxide

catalyst.54,55 The one-step process, however, incorporated all reactions in one-pot, using a

catalyst that was revealed to have both magnesia and silica.56 The process was abandoned

when catalytic cracking of oil became popular, since 1,3-BD were produced efficiently as

a byproduct of ethylene production. However, in the recent year, the shift of raw material

for ethylene production from oil to shale gas had decreased the availability of 1,3-BD,

prompting the revisiting of this old process. Improving the catalyst’s activity and stability

had been the main focus of the research, where new systems had also been established,

such as promoted MgO/SiO2,57–60 and solid Lewis acid catalyst systems.61,62

This chapter provides a review of ethanol to 1,3-BD (ETB) reaction system. This

review includes proposed reaction mechanisms, current state-of-the-art of the catalysts, and

the proposed active sites. Following the review, this chapter will discuss the techniques

that were used to study the system, as well as an outline of the thesis.

Page 31: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

11

1.1. ETB Reaction Network

1.1.1. Reaction Intermediates and Byproducts

As shown in Figure 1.1, ethanol can be converted to a variety of other chemicals.

Optimizing ethanol to 1,3-BD requires analysis of thermodynamics to select the reaction

temperature and contact time, i.e. space velocity. Byproducts that are most commonly

found during the reactions are ethylene, CO2 and acetaldehyde.63 Other byproducts that

were occasionally observed are diethyl ether, butenes, propylene, C2-C4 paraffins, diols,

acetone higher alcohols and heavy oxygenates, such as n-butanol, n-hexanol,

crotonaldehyde, and hexadienal.58,64 Dictation by thermodynamics also depends on the

catalysts employed. At different reaction temperature, acid catalysts, such as γ-Al2O3,65

will give different product distributions from transition metal catalysts, such as

CuO/SiO2.35

Optimizing the performance of the catalyst also means suppressing the formation

of unwanted byproducts, such as butenes, ethylene, and acetaldehyde. Butenes are

particularly unwanted, since at high concentration, they will form azeotrope with 1,3-BD,

which makes separation costly.60 Ethylene, on the other hand, is of much lower-value, and

its formation is followed by the production of water, which was found to poison the catalyst

by competitive adsorption.19 Another byproduct of this system is acetaldehyde, shown to

be a reactive intermediate during the reaction, which will be explained in subsequent

subchapters. Thermodynamics also determines the selectivity between acetaldehyde and

ethylene, since ethanol dehydrogenation was found to be much more endergonic reaction

than ethanol dehydration.66 Furthermore, the reactivity of acetaldehyde to undergo further

Page 32: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

12

reactions, such as polymerization, esterification, and aldol condensation, requires a careful

design of the catalyst to minimize the unwanted reactions.63,67

1.1.2. Proposed Reaction Mechanisms

Various byproducts produced during the reaction led to different reaction

mechanisms being proposed. In general, there had been five proposed reaction

mechanisms. The first two mechanisms, proposed by Lebedev4,56 and Ostromislensky68,69

were ruled out due to the improbable steps in their mechanisms. Briefly, Lebedev proposed

a reaction mechanism involving free radicals in a very complex sequence, while

Ostromislensky’s mechanism entailed reaction between ethanol and acetaldehyde, which

is followed by rearrangement of the hemiacetal to diols. Equation 1.1 shows the

stoichiometry of the reaction:

2 C2H5OH → C4H6 + 2 H2O + H2 (1.1)

Subsequently, Toussaint, et al. proposed reaction mechanism which involved

dehydrogenation of ethanol to acetaldehyde, followed by aldol condensation of two

acetaldehyde molecules, and dehydration of the aldol to give crotonaldehyde, which would

further undergo Meerwein-Ponndorf-Verley (MPV) reduction with ethanol and

dehydration to give 1,3-BD.55 Other reaction mechanisms that had been proposed were

called Prins condensation, proposed by Gruver, et al.,70 and carbanion mechanism, very

recently proposed by Chieregato, et al.71 Figure 1.2 comprehends the proposed reaction

mechanisms by Toussaint, et al., Gruver, et al. and Chieregato, et al.

Page 33: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

13

The reaction mechanism proposed by Toussaint had been touted as the generally

accepted mechanism.4,5,72,73 Aldol condensation and the subsequent dehydration steps were

reported to be facile, while the rate of ethanol dehydrogenation and MPV reduction step of

crotonaldehyde depends on the catalyst. Over MgO/SiO2 catalysts, for instance, the rate-

determining step was ethanol dehydrogenation,66 while for Lewis acid catalysts, MPV

reduction of crotonaldehyde with ethanol as the hydrogen source was thought to be the

rate-limiting step.72 There are still, however, several issues with the mechanism, including

Figure 1.2. Proposed reaction mechanisms for ethanol conversion to 1,3-BD: (a)

Toussaint’s aldol condensation; (b) Gruver’s Prins condensation; (c) Cavani’s carbanion

mechanism.

(a)

(b)

(c)

Page 34: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

14

the very rare occurrences of the reactive intermediates mentioned.71 Furthermore, acetaldol

was also shown to decompose to two acetaldehyde molecules when introduced to Ta/SiO2

catalyst.55 The very rare occurrence of the reactive intermediates was due to the rapid

reaction steps at the mostly employed reaction temperature. Second, acetaldol is a very

unstable molecule, it could either dehydrate to crotonaldehyde and water when heated, and

therefore, data analysis of the reaction should be carried out with extra caution.

The other rejected mechanism is the Prins mechanism, where ethanol undergoes

two parallel reactions, i.e. dehydrogenation to acetaldehyde and dehydration to ethylene.70

The problem with this mechanism is the ethylene protonation, an indispensable step during

the Prins mechanism, that leads to a highly unstable carbocation,4 and when this step was

not involved, the coadsorption of both ethylene and acetaldehyde on the surface is very

unstable.66 An alternative mechanism was proposed by Chieregato, et al., where ethanol

dehydrogenation resulted in two different entities.71 If the step was preceded by

dissociative adsorption, ethanol would be converted to acetaldehyde. Otherwise, a

physisorbed ethanol molecule would break a C-H bond and be converted to a stabilized

carbanion.71 This carbanion would further react with acetaldehyde to make 1,3-BD or with

ethanol to make n-butanol and 1-butene. Several issues were readily identified with this

mechanism. First issue was the use of bare MgO as the catalyst, which, if not

hydrothermally treated, would not give off 1,3-BD.71 Second, our attempts of using bare

MgO as the catalyst for Diffuse Reflectance Infrared Fourier Transformed Spectroscopy

(DRIFTS) study did not show any formation of the surface intermediates that were

demonstrated by their reports.63 Bare MgO exhibits very high absorption of CO2, which

would hamper identification of the surface intermediates that are located at similar

Page 35: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

15

wavenumber ranges. Third, the experiment was carried out under inert flow, and due to the

nature of the cascade reaction, the amount of ethanol would be insufficient to convert

further into 1,3-BD, given all the ethanol will be consumed to acetaldehyde. Furthermore,

the peak fittings were done with solely focusing on the intermediates, without considering

the possibility of the side reactions of acetaldehyde, such as acetate formation, aldol

condensation, and polymerization. Finally, the model used in the density functional study

(DFT) study did not necessarily represent the true condition of the surface. The DFT

calculation was done on a cluster MgO, with a small number of atoms used to represent

the catalyst. Stability calculation was not done either in choosing the right defect site, with

corner MgO being the model site. Later attempt by our group demonstrated that the

postulated carbanion formation does not take place when C-H bond is broken, instead, the

resulting TS further broke the C-O bond to give-off ethylene, which disputed their

proposed mechanism.66

1.2. Catalytic Systems

There had been a lot of efforts to improve the catalytic performance of the catalysts

for the synthesis of 1,3-BD from ethanol.72,73 As mentioned previously, apart from the use

of MgO/SiO2-based catalysts,57,74–77 efforts had been carried out for other classes of

catalysts, such as zirconia-based catalysts,61,78 and mixed metal oxide catalysts.79–83 The

complex reaction mechanism calls for a very demanding catalyst specification; the

catalysts have to possess balanced amount of basic, acidic, and redox sites. Typically, there

are several ways to improve the catalysts’ performance, by promotion with other metal

oxides, hydrothermal or chemical post-synthesis treatment, and modification of the

Page 36: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

16

preparation methods. This section will thoroughly discuss the current state-of-the-art of the

catalytic system.

1.2.1. Reaction Conditions and Catalytic Performance

1,3-BD had been the most realistic target molecule for ethanol upgrading, mainly

due to the abundance in ethanol supply and decline in 1,3-BD supply.60,84 Once a

competitive process, the catalytic cracking of petroleum feed had taken over the production

method, which in turn, demanded the green chemistry process to be a lot more efficient.

For the process to be competitive, there are several factors that have to be considered.

Among of these factors, 1,3-BD productivity and catalyst stability are the most important

factors. The minimum requirement for 1,3-BD productivity was suggested to be 0.15 gBD

gcat-1 h-1,83 while the catalyst needs to be stable for long hours of production, since catalyst

regeneration can also be very costly. Deactivation typically was due to coke deposition on

the catalyst, coupled with poisoning from water formed by the dehydration reaction.

Another factor that also plays important role is the reaction temperature, where

typically, the reaction takes place at 300 and 400 °C for low weight-hourly space velocity

(WHSV) of 0.2-1 h-1. The optimum reaction condition itself varied due to the nature of

different catalysts employed. For instance, unpromoted MgO/SiO2 catalysts typically

require higher operating temperature than promoted MgO/SiO2 catalysts, due to the higher

activation temperature for ethanol dehydrogenation to acetaldehyde.58,59 Other relevant

conditions are pressure and WHSV, which is essentially a unit to define catalyst to reactant

ratio. The reaction had been carried out under atmospheric pressure in a fixed bed reactor,

with the exception from Bhattacharyya, et al., who have used fluidized-bed reactor.85

Additional consideration is the ethanol to acetaldehyde ratio in reactant feed. This

Page 37: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

17

parameter is also important for one-step process, since it will be favorable to recycle some

of the acetaldehyde back to the reactor. Table 1.1 shows the catalytic performances of

recently investigated catalysts.

No Catalyst T (°C) WHSV

(h-1)

TOS

(h) X (%)

YBD

(%)

PBD (gBD

gcat-1 h-1)

Ref.

MgO/SiO2 catalysts

1 MgO hydrothermally-

treated 400 0.2 7.0 36.6 17.2 0.02 86

2 WK MgO-SiO2 (1:1) 425 1.1 4.0 52.0 17.7 0.23 58

3 MC MgO-SiO2 (1:1) 400 1.2 3.3 10.0 3.6 0.02 59

4 SG MgO-SiO2 (4:1) 400 0.4 sa N/A 53.0 14.6 N/A 87

5 WK MgO-SiO2 (1:1) 400 0.1 3.0 65.9 32.5 0.19 88

6 MC MgO-SiO2 (1:1) 400 1.0 N/A 41.2 23.6 0.14 89

7 IWI MgO-SiO2 (1:1) 300 1.1 3.3 ~3.5 ~1 0.01 60

8 WK MgO-SiO2 (13:7) 450 4.1 1.0 95.0 73.2 1.15 77

9 1% CuO/MgO-SiO2 425 1.1 4.0 74.0 36.3 0.48 58

10 2% Ag/MgO-SiO2 400 1.2 3.3 50.0 20.5 0.15 59

11 5%Ga/MgO-SiO2 (1:1) 400 0.1 3.0 98.8 52.4 0.31 88

12 2% Zn/MgO-SiO2 425 1.0 3.0 84.6 45.0 0.26 57

13 0.1% Na/MgO-SiO2 350 0.2 N/A 100.0 87.0 0.08 90

14 1.2% Zn/Talc 400 8.4 7.0 41.6 21.3 1.06 86

15 3% Au/MgO-SiO2 (1:1) 300 1.1 3.3 45.0 27.0 0.14 60

16 1.5% Zr 1% Zn/MgO-

SiO2 (1:1) 375 0.6 3.0 40.0 30.4 0.13 91

17 1.2% K/ZrZn/MgO-SiO2

(1:1) 375 0.6 3.0 26.0 27.1 0.12 91

Zr/SiO2 catalysts

1 2% Ag/4% Zr/SiO2 320 0.31 5.0 55.2 39.4 0.07 62

2 1% Ag/Zr/BEA

(Si/Zr=263) 320 0.64 3.0 30.1 19.2 0.07 61

3 3.5% Ag/Zr/BEA 320 1.2-3 3.0 - - 0.59 92

4 2000 ppm Na Zn1Zr10On 350 6.2 30.0 54.4 14.1 0.49 93

5 2% ZnO/7% La2O3/1%

ZrO2/SiO2 400 2 N/A 100.0 60.2 0.71 94

Other catalysts

1 3% Hf/9.3% Zn/HM 360 0.64 10.0 98.6 68.4 0.26 79

2 1% Cu/1% Ta/SiBEA

(Si/Al=1300) 325 0.5 3.5 87.9 63.9 0.19 95

Table 1.1 Catalytic performance of different catalysts studied for one-step ethanol to 1,3-

BD conversion. a Contact (residence) time; WHSV (weighted-hourly space velocity);

TOS (time-on-stream); X (ethanol conversion), Y (yield); P (productivity)

Page 38: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

18

1.2.2. MgO/SiO2 Catalysts

The original catalyst employed by Lebedev in his one-step process was disclosed

by Natta and Rigamonti to contain MgO and SiO2.96 MgO by itself is a very strong basic

catalyst,97,98 and its interaction with SiO2 were suggested to have formed distinct LAS,

making it a bifunctional catalyst.99 Furthermore, the defect sites of MgO are essentially a

very strong electron acceptor, making it a very strong Lewis acid-base pair.66 Proper

amount of acid-base sites was determined by changing the MgO to SiO2 ratio75,77,87,99 and

by different preparation methods.60,99,100 A contradicting result, however, was recently

reported by Baba, et al., where a bare MgO that was hydrothermally treated resulted in a

conversion of 36% and 1,3-BD selectivity of 47%.86 This MgO was reported to have a very

distinct showed that SiO2 presence was not indispensable to the catalyst.

The main issue with the catalyst system is that it is highly influenced by the Mg to

Si ratio and preparation methods. Discrepancies can be found on every considered

parameters, including the Mg to Si ratio and which preparation methods that yielded the

best performing catalysts. For instance, wet-kneading (WK),58,99 incipient-wetness

impregnation (IWI),60 mechanical mixing,59,89 sol-gel,87 and hydrothermal synthesis,86

were each reported as the best preparation method. Reports for optimum Mg to Si ratio

also contained a lot of disagreements, even for the same preparation method. For sol-gel

catalyst this parameter was reported to be between 9 to 15,87 both 1:189 and 2:159 for dry

milling, and 1:158,99 (17:3)75 for wet-kneading. In the case of promotion with transition

metals, the order of promoters impregnation and calcination also altered the activity.57,58,60

Works on unpromoted MgO/SiO2 catalysts mainly concerned the method of

preparation and oxide ratio. Wet-kneading method had been reported as the best

Page 39: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

19

preparation method.58,77,99 Wet-kneading was defined as “a process in which two or more

solid precursor materials are combined and stirred (mechanically or magnetically)

thoroughly in a liquid medium.”76,101 This method was first used by Natta in 1947,101 and

maintained its relevance through the work of Niiyama, et al.,75 Ohnishi, et al.,90 Kvisle, et

al.,99 and Angelici, et al.76 The wet-kneading was typically done with either MgO102 or

Mg(OH)258,90 as the starting material, and typically in water, since other solvent, such as

ethanol, had exhibited unwanted effect.101

Study by Weckhuysen’s group showed that wet-kneading method is superior to co-

precipitation and mechanical mixing.58 Wet-kneading method resulted in 1,3-BD yield of

~17%, significantly higher than co-precipitated catalysts (~8%) and mechanically mixed

catalyst (<1%), with 24 h time-on-stream (TOS). Co-precipitated catalyst produced

significantly higher amount of ethylene, attributed to its higher ethylene selectivity. On the

other hand, wet-kneading resulted in a layered magnesium silicate phase, which was

correlated to the higher activity of the catalyst.76 A detailed acid-base characterization

using Hammett indicator, DRIFTS with probe molecules, and TPD further produced

certain criteria for a good catalyst.100 In particular, strong basic sites have to be limited,

with participation of mostly medium and weak basic sites, in combination with some Lewis

acidity. Co-precipitated catalyst possessed combination of both strong acid-base sites,

which are shown to have increased the generation of ethylene.100

Mechanical mixing (dry milling) with different silica materials were studied by

Jannsens, et al.59 Two ordered mesoporous materials (COK-12 and MCM-41) were used

along with amorphous mesoporous SiO2. MgO was first hydrated to form Mg(OH)2, which

was followed by dry milling with the SiO2 material. The resulting dry mixture was then

Page 40: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

20

wetted with water before being dried and calcined. The wetting process appeared to have

a significant effect on the acidity-basicity of the catalyst. This process dispersed MgO,

further creating additional Lewis acid-base sites.59 Characterization of the catalyst showed

silica support covered by magnesia flakes in the case of SiO2. The other silica supports, i.e.

COK-12 and MCM-41, showed a loss of mesoporous structure when mixed with magnesia

with this preparation method. These catalysts did not show significant activity toward 1,3-

BD, with the MgO/COK-12 and MgO/MCM-41 being the worst catalyst due to their

collapsed structure. The diffusion of MgO into the lattice was suggested to limit access of

reactant to the active sites, which was also aggravated by the presence of large magnesium

silicate phases.59

Sol-gel was another explored preparation method.87 Starting from Mg/Si oxide ratio

of 2 up to 9, significant fosterite, a magnesium silicate, phase was identified using X-ray

diffraction (XRD) and attenuated total reflectance infrared spectroscopy (ATR-IR). This

fosterite phase, however, was detrimental to the process, since it possessed more acid sites,

leading to an enhanced ethylene yield. High amount of Mg, i.e. Mg/Si > 15 accumulated

the alkenol intermediate, due to the limited amount of acid sites. The best performing

catalysts exhibited combination of a limited number medium-strength acid sites with strong

basic properties, which contradicted Weckhuysen’s finding.87,100 The LAS, however, was

observed to be transformed into BAS when water was formed as the byproduct.

MgO/SiO2 with various MgO loading was synthesized using IWI method.60 For

loading between 10-80%, a magnesium silicate hydrate (MSH) phase was observed at

~50% loading. Higher loading at 80% showed a crystalline MgO phase, presumably

formed by excessive MgO formation covering SiO2. 29Si magic-angle spinning (MAS)

Page 41: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

21

NMR spectroscopy showed that at lower loading at 10 and 30% MgO, SiO2 still retaine its

support characteristics, where Mg2+ cations mainly interact with the isolated silanol groups.

The Q3 feature was caused by silanol groups [Si*(OSi)3(OH)] and [Si*(OSi)4]. However,

at higher loading, new peaks associated with [Si*(OMg)(OSi)3], [Si*(OMg)(OSi)2(OH)],

and [Si* (OMg)2(OSi)2] were formed. As expected, higher MgO loading increased the

basicity of the catalyst, demonstrated by CO2 DRIFTS. The reactivity study showed that

higher loading significantly reduced the catalyst’s conversion while increasing selectivity

toward both acetaldehyde and 1,3-BD. The high silanol content for lower loading was

suggested to be responsible for ethanol dehydration at lower loading.

Other unpromoted MgO/SiO2 systems that have been studied are clay materials and

talc.86,103,104 Without promotion with transition metal oxides, clay, a naturally occurring

magnesium silicate mineral, favored dehydration of ethanol to produce ethylene.103,104 Talc

is a 1:2 layered structure, where a unit cell includes six octahedral sites and eight tetrahedral

sites. Mg2+ ions represent the former, while cations for the latter are Si4+. Significant

improvement of these magnesium silicate materials was attained by promotion with

different transition metal catalysts. Nickel,104 manganese,103 and zinc86,105 were all

incorporated into the catalyst to achieve higher conversion. Characterization of Zn/talc

carried out using ICP, XRD, and XPS suggested that Zn was incorporated into the lattice.

The Zn site in the catalyst was shown to be responsible for the enhanced ethanol

dehydrogenation step, which increases the overall 1,3-BD yield.86

Promotion using other metal oxides were very commonly done to improve 1,3-BD

activity. Mainly, this was done to add redox sites to the catalysts and hence, to reduce the

energetic barrier of the rate-limiting step. Larina, et al. published a study on the role of

Page 42: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

22

ZnO as a promoter for magnesia/silica catalyst on the ethanol conversion to 1,3-BD.57 The

amount of both Lewis and BAS were investigated using pyridine adsorption spectroscopy.

The experiment revealed that there are two types of LAS in the catalyst, one in MgO-SiO2

contact phase and one in ZnO-SiO2. When tested over MgO single oxide, the formation of

1-butanol was preferred, indicating the necessity of both acidic and basic sites in the

catalysts. However, the excessive silica content would lead to the preferential formation of

ethylene, since the number of acid sites would significantly be increased. Catalytic testing

conducted demonstrated the enhanced yield of 1,3-BD attributed to the ZnO role in

catalyzing dehydrogenation of ethanol as the first step of the reaction. The author

postulated this step to be the rate-determining step for the complete mechanism 57.

Other transition metal oxides have also been used in the past to improve the

MgO/SiO2 catalysts’ performance. Angelici thoroughly studied 1%CuO/MgO/SiO2

catalysts, where they saw an enhanced improvement, with Cu0 as the active sites suggested

by operando X-ray absorption near edge structure (XANES) spectroscopy.106 Deposition-

precipitation (DP) of Au onto MgO/SiO2 catalyst also saw significant increase in the

MgO/SiO2 catalyst.60 DP of AuCl-based precursor surprisingly completely transformed the

bulk MgO into MSH phase. This phenomenon was attributed to the Cl- effect from the Cl-

ion produced from the precursor’s hydrolysis, since acetate precursor did not transform the

MgO into MSH. In-situ titration experiment with propionic acid, a very weak acid, showed

that 1,3-BD production considerably decreased, and did not recover when propionic acid

was not fed anymore. The dehydrogenation reaction to acetaldehyde recovered, however,

indicating the presence of weak basic sites to catalyze dehydrogenation and strong basic

sites to catalyze the subsequent reactions.60 Another transition metals that were used to

Page 43: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

23

improve 1,3-BD yield is Ag.59 Silver was dispersed on silica phase when impregnated over

calcined MgO/SiO2 catalyst, as shown by ambient SEM and EDS analysis. Synergistic

effect was observed, where Ag catalyzed the dehydrogenation of ethanol, while MgO-SiO2

phase catalyzed the subsequent reactions. However, increase in Ag loadings seemed to

have created aggregated, larger Ag particles, which in turn deactivated the catalyst.59

Other types of metals had been used as well as promoters. Ohnishi, et al., for

instance, prepared three MgO-SiO2-Na2O and MgO-SiO2-K2O by impregnating aqueous

NaOH and KOH with MgO-SiO2 catalysts prepared from different raw materials using

different methods.90 They found out that the (0.01%) Na2O promoted magnesia-silica

catalyst prepared from ethyl orthosilicate and magnesium nitrate (1:1) by wet-kneading

produced the highest catalytic activity (100%) and selectivity (87%) for formation of 1,3-

BD at 350 ºC for the one-step process.90 The origin of this alkali metal promotion, however,

is not known since there was no characterization reported on how Na and K is coordinated

to the surface. Zirconia, another class of catalyst that will be discussed below, is another

non-redox promoting metal that was used to improve 1,3-BD formation.91 Jones, et al.

found that IWI over an uncalcined MgO/SiO2 gives a higher surface area, since there were

more OH groups to interact with the promoter. Zn itself improved the dehydrogenation

reaction but lagged the aldol condensation, demonstrated by accumulation of acetaldehyde

in the product stream. Co-promotion with zirconia, another solid Lewis acid, was shown

to significantly improve this, since it further provided additional aldol sites on top of the

support’s native sites. The authors suggested that both ZnO and ZrOx were more readily

dispersed over Mg-O-Si linkages, and coprecipitation method was shown to generate more

of this linkage, as shown by 29Si MAS NMR. Furthermore, CHCl3 and NH3 DRIFTS

Page 44: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

24

experiments suggested that post-treatment of the catalyst with alkali metals, such as Na,

Li, and K, poisoned the stronger acid sites, without introducing strong basic sites. This

poisoning effect significantly removed ethylene from byproduct, and therefore opening

possibilities of recycling the product back to the system.91

1.2.3. ZrOx –based Catalysts

Study on ethanol to 1,3-BD over ZrOx-based catalysts had been extensively studied

recently. Toussaint, et al.,55 Jones, et al.,78 and Ivanova, et al.,62 screened a number of

different Lewis acid catalysts that would work on aldol condensation and MPV reduction.

The Lewis acid material can then be combined with redox metal catalyst that would

improve the conversion of the dehydrogenation of ethanol. Jones, et al. suggested the use

of ZnO and ZrOx,78 while Ivanova, et al. identified AgO/ZrOx as the most active catalyst.62

The complete reaction network for the catalytic conversion of ethanol into 1,3-BD

over metal-containing (M=Ag, Cu, Ni) oxide catalysts (MOx=MgO, ZrO2, Nb2O5, TiO2,

Al2O3) supported on silica was investigated by Sushkevich, et al.62 From the reaction

network above, these authors narrowed down their catalyst selection to target the

dehydrogenation of ethanol, aldol condensation of acetaldehyde, and MPV-reduction of

the crotonaldehyde. Their experiments demonstrated the superior Ag and Cu-promoted

catalysts over Ni-promoted. Nb and Al-based oxide catalysts developed higher selectivity

toward ethylene due to the BAS, while magnesia and titania oxide catalysts were prominent

in catalyzing aldol condensation. The best performing catalyst was found out to be 1%

Ag/10%ZrO2 on silica reaching 88% conversion and 74% yield of 1,3-BD at 593 K, with

WHSV of 0.04 g g-1 h-1 and time on stream (TOS) of 5 hours.62

Page 45: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

25

Optimization of the Ag/ZrOx catalysts was subsequently done by fine-tuning the

composition of both Zn and Zr.61 Catalysts employed were ZrBEA zeolites with various

Zr/Si ratios and zirconia supported on silica, all promoted with silver. Characterization of

the acid sites of the catalysts was conducted by FTIR spectra of deuterated acetonitrile. The

silanol group of the catalysts contributed to the high number or the BAS, giving peaks at

2275 cm-1 that readily disappeared upon evacuation, pointing to the weak interaction of the

OH group. The LAS were also determined using the same technique, particularly by

looking at the band centered at 2303 cm-1. The number of LAS apparently played

significant role in determining which reaction pathways to take place; catalyst with the

highest yield turned out to be the one with the highest LAS. The best catalyst performance

was observed for Ag-promoted ZrBEA (Si/Zr = 100), with 1,3-BD selectivity of 56% and

conversion of 48% under the following conditions: T = 593 K, WHSV = 0.32 g g−1 h−1,

observed after TOS = 3 h.61

DFT calculation, in combination with CO-DRIFTS study, elucidated the nature of

the LAS on the catalyst.107 Two active sites, open and closed isolated LAS, were shown to

be available on the surface. The open site, where there is one terminal hydroxyl group

coordinated to Zr (HO-Zr-(OSi)3), was found to be the main catalytic active sites.107 Over

series of synthesized catalyst, direct correlation was made between the 1,3-BD productivity

and the relative amount of open LAS. The reaction mechanism over this catalyst was

further probed using combination of SSITKA and isotope labeling.108 As expected, two

sites were available for the reaction, the silver sites that were responsible for ethanol

dehydrogenation, and Zr LAS that catalyzed the subsequent aldol and MPV reactions.

Deuterated ethanol CH3CD2OH and CH3CH2OD showed that ethanol dissociatively

Page 46: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

26

adsorbed on silanol site, followed by C-H bond breaking over Ag/Si-OH site. The resulting

acetaldehyde was then desorbed from the surface.109 Enolization of acetaldehyde was made

possible by Zr LAS and Zr-O pair sites, and reaction with vapor-phase acetaldehyde

molecule was carried out according to Eley-Rideal mechanism. Facile dehydration of

acetaldol then followed, where crotonaldehyde was produced. MPV reduction by ethanol

was then determined to follow the Langmuir-Hinshelwood mechanism by fitting of kinetic

data.108

Combination of Zn and Zr for ETB reaction had also been explored.93,110 Similar to

previously published endeavors, Zn/Zr mixed oxides were used. From IR-pyridine probing

experiment, it was found that the acidity of the catalyst as a function of Zn/Zr ratio had a

maximum at Zn/Zr = 1:10. The use of Na as a promoter to control the surface acid-base

properties was confirmed by NH3-TPD and IR-Py as well. When 2000 ppm if Na was used

as a promoter, the balanced basic sites and weak BAS exhibited selectivity of 47% at 97%

conversion, with higher productivity per grams of catalyst. Another alkali metal oxide,

cesium oxide, was studied as well as a promoter that eliminates acid sites and reduces

ethylene formation.110 The similarity between Zn/Zr and Ag/Zr catalysts suggested similar

reaction mechanism and active sites. Zn/Zr catalysts, however, required a third promoter

in the form of very small amount of alkali metal to eliminate the acidity it provided, unlike

Ag/Zr catalyst, which did not provide improved Lewis acidity.107

1.2.4. Other Catalysts

This subsection will discuss other metal oxide catalysts that had been

used/investigated in the past. Discussion will cover some of the catalysts that have similar

Page 47: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

27

characteristics with Zn(Ag)/Zr catalyst, i.e. Lewis acid combined with dehydrogenation

catalysts, such Hf-based catalyst79 and Ta-based catalyst.82,95 Due to the extensive work

carried out by Sushkevich and Ivanova, Zr-based catalysts were classified as a different

class of catalyst.43,61,62,107–109

Bhattacharyya and Ganguly performed one-step catalytic conversion of ethanol to

1,3-BD in a fixed bed reactor on various single oxide catalysts of aluminum, thorium,

magnesium, iron (III), and zirconium.111 The maximum process yield of 1,3-BD was

achieved at 36.1% using thorium oxide prepared from thorium nitrate and ammonia

carbonate with flow rate at 1.256 mL/g h-1 at 450 ºC. All metal oxides were selected based

on their capabilities of catalyzing both dehydration and dehydrogenation reactions.111 Few

years later, Bhattacharyya and Avasthi conducted exhaustive experiments on the

conversion of ethanol to 1,3-BD via the one-step process on single alumina oxide and

binary alumina-zinc/calcium/chromium/magnesium oxide catalysts. They reported

maximum yield of 72.8% on a fluidized bed reactor using the binary alumina-zinc oxide

catalyst (60:40) as opposed to 55.8% yield on a fixed bed reactor at 425 ºC.85

Work on hafnium oxide catalyst, promoted with Cu and Zn as dehydrogenation

promoter, was done by Baerdemaeker, et al.79 Building on Jones’ catalyst screening,78

slight modification was performed to understand which element contribute to which step

of the mechanisms. Switching the precursors from nitrate to chloride already contributed

in higher 1,3-BD selectivity, stability, and lower ethylene selectivity. Replacing zirconium

with hafnium, a softer material but still in the same group as Zr, suppressed the formation

of ethylene more significantly. Their study also revealed Zn’s superiority in catalyzing

dehydrogenation of ethanol, since when Zn single oxide was used, acetaldehyde

Page 48: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

28

accumulated without significant formation of 1,3-BD. It was inferred that, the addition of

Hf suppressed ethanol dehydration, while also promoted the C-C coupling step, i.e. aldol

condensation.

Tantalum oxide, a rare transition metal material, was among the first catalysts used

to catalyze the reaction in the two-step process.68,81 Its use had been revisited following the

renewed interest in the Lebedev process.82,95,112–114 Kyriienko, et al., synthesized a

TaSiBEA zeolite material by dealuminating the starting AlSiBEA with HNO3, followed by

the introduction of tantalum into the framework. 95 The tantalum itself was present as

isolated mononuclear in the framework, as confirmed by UV-Vis and XRD. The catalyst

was further promoted with Ag, Cu, or Zn to introduce dehydrogenation sites. UV-Vis

characterization suggested that Ag was present as Ag(I) and oxidized silver cluster, Cu was

available as isolated mononuclear and oxidized cluster, while Zn was suggested to be in

the framework and as a polynuclear zinc oxide in the extra-framework position.95

Incorporation of Cu and Zn is suggested by pyridine DRIFTS to give higher amount of

LAS, as compared to Ag promotion, while ZnO incorporation exhibited higher BAS

content in the catalyst than the rest. Further deuterated chloroform DRIFTS also suggested

that incorporation of the transition metals into the TaSiBEA catalysts had eliminated

medium strength basic sites, replacing them with weaker ones.95 Mechanistic study for Ta-

SiBEA catalyst,82 and subsequently on Ag-Zr-BEA,114 was done by Müller, et al.82,114

Operando modulated DRIFTS-MS was used to control the ratio of ethanol to acetaldehyde,

maintaining sufficient ethoxy coverage to keep replenishing acetaldehyde formation by

both dehydrogenation and MPV reduction.114 The presence of Ag was reduced

nanoparticles enabled reduction of crotonaldehyde in two possible mechanisms, direct

Page 49: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

29

hydrogenation by hydrogen molecule under low ethanol pressure, and MPV reduction

under high ethanol pressure.114

2. Approach

2.1. Approach

The recent emergence of combined dehydrogenating/Lewis acid catalysts has

opened an alternative for the Lebedev’s process. Despite this, the traditional MgO/SiO2

catalysts are still very interesting, due to the lack of suitable characterization methods. Mg

to Si ratio and other synthesis parameters for the available preparation methods are the

most enticing part of the research. Additionally, the catalytic system is still not well-

understood, despite attempts to characterize the system. This is mainly due to the nature of

the catalyst itself that limits the use of several in-situ (operando) spectroscopic methods.

In this study, MgO/SiO2 catalysts were used as the investigated catalysts, where reaction

mechanisms, role of promoters, and catalyst characterization are integral parts of this

dissertation. The approach taken to achieve these objectives consisted of the following

steps:

1. Perform DFT calculation as a preliminary study on proposed reaction mechanisms over

an ideal model MgO surface.

2. Synthesize MgO/SiO2 catalyst with wet-kneading method, fixing Mg to Si ratio to 1

with varying calcination temperature at 500 and 800 °C.

3. Perform reactivity study and reaction mechanism study, both surface and vapor-phase

intermediates.

Page 50: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

30

4. Intensively characterize the catalyst using probe molecules and in-situ spectroscopy,

correlate the catalytic active sites with activity and selectivity.

5. Promote the MgO/SiO2 catalyst with transition metal (Cu and Zn), characterize the

system, and investigate the catalytic consequences.

DFT calculation was performed using VASP software package, with MgO (100)

surface termination chosen as the model catalyst. Specifically, a kink site, Mg2+3C O

2-4C pair

were chosen as the active site, since defects had been suggested to have a very high

catalytic activities.115 This first approximation simplified the current state-of-the-art, where

Lewis acidity and basicity were represented by the lower coordinated ionic pairs on the

surface. In reality, the Lewis acidity and basicity were suggested to be provided by the Mg-

O-Si linkages, and promotional effect on the redox site was provided by transition metal

sites on the catalyst.

Experimentally, MgO/SiO2 catalysts were synthesized by wet-kneading in water.

Mg(OH)2 from hydrothermal-NaOH assisted precipitation of magnesium nitrate precursor

and Cab-O-Sil EH5 fumed silica were used as a starting material. Wet-kneading was done

for 4 hours, before separation, drying, and impregnation/calcination. All final catalysts

were calcined at elevated temperature to transform the material into oxides. Experimental

study comprised in-situ and operando spectroscopy, kinetic experiment using fixed bed,

and other bulk and surface ambient characterization such as TEM, SEM, XPS, and XRD.

Materials used were ethanol in inert, such as N2, He, and Ar. Probe molecules were used

to investigate the reaction mechanism, acidity, and basicity of the catalyst.

Page 51: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

31

2.2. DFT Calculation

Theoretical (computational) chemistry methods provided an alternative to predict

reaction mechanisms, thermodynamics, kinetics, and even catalyst design by creating

potential energy landscape. The potential energy surface (PES) was created by

optimization of different minima and maxima (TS). This iterative step is essentially solving

the Schrödinger’s equation,

ĤΨ(R,r)=EΨ(R,r) (1.2)

The Hamiltonian in this equation can be separated into two terms, one to account for the

electron and the other for the nucleus. This separation is made possible by using the Born–

Oppenheimer approximation.116 Consequently, the wavefunction is represented as a

product of the electronic and nuclear wavefunctions. Further simplification is also made

by ignoring the nuclear kinetic energy and only taking into account the nuclear’s potential

energy and electron’s kinetic energy. Ab-initio methods and DFT study aim to solve the

Schrödinger’s equation by using electronic wavefunctions in the forms of single electron

molecular orbitals.116

DFT is less computationally expensive than post-HF calculations. It uses electron

(probability) density, ρ(r) to compute the electronic energy. A functional is used to

represent the energy, because energy is a function of electron density, which is a function

of position. This electron density is represented by a sum of one electron orbitals in the

Kohn-Sham equation, 𝜌 = ∑ 𝜑𝑖(𝑟)𝑁𝑖 . The DFT energy is further calculated in the Kohn-

Sham equation,

𝐸[𝜌] = 𝑇𝑠[𝜌] + ∫ 𝑑𝒓 𝑣𝑒𝑥𝑡(𝒓)𝜌(𝒓) + 𝐸𝐻[𝜌] + 𝐸𝑋𝐶[𝜌] (1.3)

Page 52: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

32

The terms on the right-hand side represent the (electronic) kinetic energy, external potential

acting on the interacting system, Hartree (or Coulomb) repulsion energy, and the exchange-

correlation energy.116 The exchange-correlation energy is approximated, and several

approximations have been developed, such as the local density approximation (LDA), local

spin density approximation (LSDA), and generalized gradient approximations (GGA). The

generalized gradient approximation (GGA) is an adaptation of the LDA that accounts for

inhomogeneity of the electron density, in which the non-local correction of the gradient of

the electron density (moving away from the coordinate) is added to the exchange-

correlation energy.116 In this work, GGA was used to approximate the exchange-correlation

energy, specifically the simplified Perdew-Burke-Ernzerhof (PBE) GGA. More details of

the computational calculation are provided in Chapter 2

2.3. In-situ and Operando Spectroscopy

In-situ spectroscopy is defined as spectroscopic characterization under relevant,

operating condition.117 The dynamics of the reaction and catalysts at different temperature

can be studied by manipulating the temperature and pressure in a controlled manner.

Operando spectroscopy, on the other hand, is essentially in-situ spectroscopy with

monitoring of the vapor-phase product identification.118,119 Typical setup for operando

spectroscopy comprises a reaction cell with a temperature and/or pressure controller. The

use of this reaction cell enables probing using optical light source, such as Raman and FTIR

spectroscopy, at reaction condition, with flowing gas. The output of this reaction cell is

then connected to product identification system, either GC-MS or MS. Figure 1.3 shows a

typical setup for operando spectroscopy.

Page 53: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

33

2.3.1. Infrared Spectroscopy

Infrared (IR) spectroscopy is based on the radiation absorption by the characterized

compounds. When an IR beam is passed through a sample, the compound absorbs radiation

at a specific wavelength, resulting in a decrease of the transmitted radiation. In the spectra,

the absorption is reflected as a dip in transmission, or a peak in absorbance. Upon absorbing

the IR radiation, the molecules’ oscillation amplitude will increase, hence the molecule is

excited to a higher vibrational state. Vibrational states are quantized energy levels, and the

particular wavelength of absorption by a specific bond depends on the energy difference

between the ground level and the excited state. Hence, different bonds produce different

peaks/dips, which also vary depending on the different oscillation modes, in the IR spectra

at different wavenumbers.

Monitoring the molecular events taking place at a certain reaction temperature or

pressure, i.e. in situ IR spectroscopy, is now possible and widely used. A substrate, or

Figure 1.3. Operando spectroscopy setup, flow reaction cell temperature/pressure

controller equipped with FTIR, UV-Vis and Raman spectroscopy that enables real-time

online measurement. Output is connected to real-time GC/MS system. Adapted from: http://www.lehigh.edu/operando

Page 54: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

34

catalyst, can be put into a reaction chamber, where a reactant is flown over it. The

spectroscopy is then used to continuously monitor the surface species present on the

substrate while the surface is heated up to a certain desired temperature. This

characterization method is very powerful because one can get information on the bonds

created or destroyed during the temperature ramping, which, when properly analyzed,

could give information on the possible transition states during the reaction.

Vibration spectroscopic measurements of the solid/gas interface can be effectively

carried out using reaction chamber equipped with optical windows enabling measurements

in transmission mode, attenuated total reflectance (ATR) and total or diffuse reflectance

modes. Fourier transform infrared spectroscopy (FTIR) is generally limited to powdered

(submicron-sized) materials, given limitations in throughput.

2.3.2. UV-Vis-NIR Spectroscopy

UV-Vis-NIR spectroscopy is another powerful tool to study catalysts. UV, Vis

(visible), NIR (near infrared) regions cover 200-400 nm, 400-800 nm, and 800-2500 nm

wavelength, respectively.120,121 The UV-Vis region is very useful since it probes the

electronic transitions, while NIR can discover overtones and combination bands of

fundamental vibrational vibrations. This spectroscopy technique is especially useful when

studying the transition metal ions, such as Cu and Zn in this dissertation study. In particular,

two types of bands can be identified in the spectra, charge transfer (CT) bands and d-d

transition bands. CT bands are usually associated with the highest oxidation state, while d-

d transition bands represent reduced states. Bands associated with plasmon resonance are

also available at higher wavelength region, which signifies the presence of a metallic state.

Page 55: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

35

Coordination number of the transition metal species can be determined as well

using UV-Vis spectroscopy.122 This is done by comparing the optical absorption edge

energies or band gaps in metal compounds with properties resembling crystalline or

amorphous semiconductors. Several references with known coordination numbers, i.e.

isolated, dimer, trimer, are tested and plotted against the band gap. The band gaps

themselves are obtained from the hν-intercept value [(αhν)1/η = 0] by extrapolating a

straight line from the linear region near the edge on the Tauc plot, where α is absorption

coefficient, hν is the energy of the incident photon, E0 is the optical absorption edge energy,

and for crystalline semiconductors, η is 0.5, 1, 2, and 3, when the optical transitions caused

by photon absorption are direct-allowed, direct-forbidden, indirect-allowed, and indirect-

forbidden, respectively, whereas for amorphous, homogeneous semiconductors, η is

typically 2.122

2.3.3. Operando XANES and EXAFS

High energy X-ray spectroscopy comprises X-ray absorption near edge

spectroscopy (XANES) and extended X-ray absorption fine structure (EXAFS). Sample is

irradiated with a tunable source of high intensity monochromatic x-rays from a synchrotron

radiation facility. The bulk of the solid is penetrated by the x-ray, and only materials with

nearly all the atoms on the surface give surface information. Oxidation state and

coordination information can be extracted using this characterization method. The

characterization can also be done in operando mode, which allows correlation with

catalytic activity.

Not only does XANES give information for oxidation states, but it also

discriminates tetrahedral and octahedral coordination of metal oxides by identification of

Page 56: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

36

the pre-edge features. Octahedral coordination has a center of symmetry (unlike

tetrahedral), and transitions between states of g symmetry are not dipole allowed. Hence,

octahedral coordination will not exhibit a pre-edge feature. EXAFS, on the other hand,

provides information on the bond length and coordination number of different atoms on

the molecules. With the correct reference, interpretation can be made to correlate the

change in the bond length, oxidation states, coordination number, or change of coordination

type during the reaction.

2.3.4. Temperature-Programmed Reaction Spectroscopy

Temperature-Programmed reaction spectroscopy (TPRS), often called TPD

(temperature-programmed desorption), TDS (thermal desorption spectroscopy), and TPSR

(temperature-programmed surface reaction), is a very powerful method to characterize a

catalyst.123 When only desorption is taking place, it is either called TDS or TPD, and when

there is a reaction involved, i.e. decomposition, reduction, oxidation, it is called TPRS or

TPSR. This technique requires a very simple infrastructure; a tube with a furnace

containing the catalyst sample and mass spectrometer (MS) that is equipped with vacuum

pump. Typically, the heating rate applied is constant, and hence, partial pressure of the

vapor-phase can be related to the desorption (reaction) temperature. Redhead had shown

that for fast pump rate, the desorption rate is proportional to the pressure in the chamber.124

The desorption kinetics can also be determined, with the desorption activation energy

estimated from the peak temperature. For the first-order kinetic, the following equation is

applicable:

𝐸𝑎

𝑅𝑇𝑝2 =

𝜐

𝛽𝑒

(−𝐸𝑎

𝑅𝑇𝑝) (1.4)

Page 57: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

37

In this study, this method was used to identify the intermediates desorbed from the

surface, further confirming the reaction mechanism and rate-determining steps. Ethanol or

other reactants were adsorbed on the surface, flushed to remove the physisorbed species,

and finally temperature was ramped under reactant flow or under inert flow. Depending on

the reaction mechanism, activation energy can be determined, as discussed by Redhead.

Deconvolution of the curve of a m/z number is particularly important. A m/z number can

represent more than one molecules. For instance, m/z = 31 could originate from 1,3-

butanediol, n-butanol, 2-butanol, crotyl alcohol, ethanol, and diethyl ether. Therefore, it is

very important to deconvolute the curve and to consider the important m/z and discard the

others.

2.4. High Sensitivity Low Energy Ion Scattering Spectroscopy (HS-LEIS)

High Sensitivity-Low Energy Ion Scattering (HS-LEIS), was used to probe the

outermost layer of the catalyst, since reaction takes place on the catalyst surface.

Quantitative, elemental information of the outermost layer is probed by bombarding the

surface perpendicularly with noble gas ions. The noble gas ions, which are lighter than the

elements on the surface, will be backscattered, recorded, and analyzed based on the classic

laws of mechanics (conservation of momentum and conservation of energy).125 The

equation can be seen below:

E𝑓 = k2 (𝑚2

𝑚1θ) 𝐸𝑖 (1.5)

Ei is the initial energy carried by the noble gas ion, m1, θ is the backscatter angle, m2 is

the scattering surface atom, k is a known function of m2/m1 and θ, and Ef is the final energy

of the scattered atom that is analyzed by the analyzer. The m2 can hence be determined

from the equation and the element can be determined.125 Elemental composition calculation

Page 58: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

38

can further be estimated using quantum chemical method, by first deconvoluting the curve

using Gaussian fitting and integrating the area under the curve to get Ii and use the

following formula:

C𝑖 =

𝐼𝑖

√𝑚𝑖⁄

∑ 𝐼𝑖

√𝑚𝑖⁄

(1.6)

2.5. Probe Molecules

Characterization of (supported) metal oxide catalysts often involves the use of

probe molecules. Probe molecules, combined with spectroscopic techniques, such as

DRIFTS and TPD, can provide a lot of insights on the catalysts surface. This

characterization technique can be conveniently done at relevant reaction temperatures,

which further provides the most relevant information regarding the chemical nature of the

catalyst. In the case of bifunctional catalysts such as MgO/SiO2 catalysts, acidic and basic

probe molecules are indispensable for the acid-base characterization. Two probe molecules

are typically used to characterize the acid sites of a catalysts: ammonia and pyridine.

Ammonia is the most common probe molecule, since it is a very small molecule with

relatively strong basic characteristic, allowing it to penetrate the small pore of the catalysts.

Furthermore, ammonia adsorbs on both BAS as ammonium ion NH4+

and as NH3 by

donating its excess electron pairs to LAS. DRIFTS spectroscopy can be used to

qualitatively and semi-quantitatively determine the acidity of the catalyst. Pyridine, on the

other hand, is a weak but bigger basic molecule. This weak base can differentiate between

LAS and BAS, and is more sensitive to the strong acid sites.

The basicity of the catalysts were probed with various acid probe molecules. CO2

remains the most popular probe molecules, due to its slightly acidic nature and its

Page 59: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

39

versatility to bond with basic sites with different strengths. Monodentate, bidentate,

symmetrical, and polydentate (bridged) carbonates are readily formed on the surface with

different splitting values, while bicarbonate can also help identify the presence of weaker

basic sites.126 Methanol can also be used as a reactive probe molecule. On metal oxides,

methanol adsorbs both as a Lewis-bound molecule and as a methoxy species. Upon thermal

treatment, these two species desorb as both molecular methanol and as its reaction product,

depending on which sites it is adsorbed to. On acid sites, methanol undergoes dehydration

to yield dimethyl ether, while on basic sites, CO2 will be produced by consecutive C-H

bonds breaking. Redox site, when available, will transform methanol into formaldehyde.

This versatility is very important when considering a catalyst that possesses different

functional active sites.

2.6. Product Determination with GC-MS/FID

Gas chromatography (GC)-mass spectrometry (MS)/flame ionization detector

(FID) is a very important tool to identify volatile chemical products. The setup essentially

consists as a gas chromatography, coupled with MS or FID as the detector for product

analysis. A GC is an oven box containing capillary or packed column that is internally

coated with a polar solute, which is defines as stationary phase. The oven box will

controllably heat up the column for separation purpose. On top of the oven, there is a

sample loop that functions as a sample storage. At one state, the GC valve will allow sample

to be continuously flown to the sample loop for a storing purpose. At another state, the GC

valve will allow carrier gas, or mobile phase, to flush the sample loop into the GC column

for product identification purpose.

Page 60: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

40

When the carrier gas introduces sample to the GC, molecules will be adsorbed on

the stationary phase, while the mobile phase keep flowing through the column. Separation

is based on two variables, the affinity of the molecules (based on polarity) to the column

and boiling point of each molecules. Separation is aided by the heating oven which will

desorb each molecule based on their boiling point. Separation of molecules with similar

boiling points will be based on the affinity to the column. The separated, desorbed

molecules will be carried to the detector, MS and/or FID. Mass spectrometry detects

volatile products by ionizing them and analyzing their mass-to-charge ratio, which is very

specific to different molecules. The ionized molecules are then accelerated to the same

kinetic energy by charged plates down the MS. The ions are deflected using a magnetic

field selectively allowing one m/z to hit the detector. Simultaneous product identification

is made possible by rapidly changing the magnetic field. However, MS is not inherently

quantitative.127

For quantification purpose, FID detector was used in this work. FID detects the ions

formed during combustion of organic compounds in a hydrogen flame.126 These ions are

proportional to the concentration of organic species in the sample. Since this detection

method is based on combustion, the detectable molecules are limited to organic molecules

with C-H bonds. Unlike MS, where the peaks can be identified as a specific molecule, there

is no qualitative way of identifying a single molecule without the use of a standard

molecule. Standard molecules can be used to determine the order of the GC retention time,

and then calibrated to the area recorded by the GC-FID, and response factor for each

molecule can be determined as the area/concentration.

Page 61: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

41

3. Thesis Outline

Chapter 2

The complete experimental methods were presented in this chapter, which covers

all experiments/computation done in this work. Results and discussions were presented in

the following chapter accordingly.

Chapter 3

This chapter discusses the study of relevant reaction mechanisms by DFT

simulation. Three previously proposed reaction mechanisms were investigated over a kink

site of MgO (100) crystal model catalyst. This study was represented by comparison of

kinetics and thermodynamics of the three reaction mechanisms, demonstrated by plotted

potential energy surfaces. This work was published in J. Catal. 2017, 346, 78–91.

Chapter 4

This chapter follows up the mechanistic study using in-situ DRIFTS to investigate

the surface reaction mechanism on the surface of MgO/SiO2 catalyst. Detailed study using

different feed, such as crotonaldehyde/ethanol, acetaldehyde, ethanol/inert, and

ethanol/ethanol, were used in combination with GC-MS to identify the product stream.

This work was published in Catal. Sci. Technol. 2017, 7 (20), 4648–4668.

Chapter 5

This chapter discusses the catalyst characterization of MgO/SiO2, as well as the

kinetics of the reaction. Identification of the active sites were done by titration of the

catalyst, as studied using DRIFTS and fixed-bed reactor with GC-MS. Characterization

was done using XRD, HS-LEIS, and probe molecules. This work was submitted to Journal

of Physical Chemistry C journal.

Page 62: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

42

Chapter 6

The catalyst was promoted with transition metal, i.e. Zn and Cu. Initial

computational study using a first approximation for the catalyst model was done to

compare the reaction mechanisms on promoted catalysts. Promotional effect was briefly

discussed by comparison of reactivity data, which is followed by characterization using

UV-Vis spectroscopy, TEM and DRIFTS of CO2 and NH3. Subsequently, the chemical

nature of the catalyst was studied by operando methanol DRIFTS to probe the chemical

change of the catalyst brought about by promoting the catalyst with transition metals. The

catalysts were further studied for its reaction. In-situ DRIFTS and cofeeding with different

poison probe molecules were carried out to study the active sites of the catalyst. This study

was completed by operando XANES/EXAFS to investigate the chemical change of the

promoter materials during reaction. Part of this work was submitted to ACS Catalysis

journal.

Chapter 7

A summary of the whole work for this dissertation is presented. Future outlook was

discussed as well, presenting works that need to be further done.

Page 63: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

43

References

(1) McCoy, M.; Reisch, M. S.; Tullo, A. H.; Short, P. L.; Tremblay, J. F.; Storck, W.

J. Chem. Eng. News 2007, 85, 29.

(2) Werpy, T.; Petersen, G. Top Value Added Chemicals from Biomass. Volume 1.

Results of Screening for Potential Candidates from Sugars and Synthesis Gas; US

Department of Energy: Washington DC, 2004.

(3) Keturakis, C. Operando Molecular Spectroscopy during Catalytic Biomass

Pyrolysis, Lehigh University, 2015.

(4) Angelici, C.; Weckhuysen, B. M.; Bruijnincx, P. C. a. ChemSusChem 2013, 6 (9),

1595–1614.

(5) Sun, J.; Wang, Y. ACS Catal. 2014, 4 (4), 1078–1090.

(6) Ni, M.; Leung, D. Y. C.; Leung, M. K. H. Int. J. Hydrogen Energy 2007, 32 (15),

3238–3247.

(7) Haryanto, A.; Fernando, S.; Murali, N.; Adhikari, S. Energy & Fuels 2005, 19 (5),

2098–2106.

(8) Hanspal, S.; Young, Z. D.; Shou, H.; Davis, R. J. ACS Catal. 2015, 1737–1746.

(9) Morrill, M. R.; Thao, N. T.; Shou, H.; Davis, R. J.; Barton, D. G.; Ferrari, D.;

Agrawal, P. K.; Jones, C. W. ACS Catal. 2013, 3 (7), 1665–1675.

(10) Birky, T. W.; Kozlowski, J. T.; Davis, R. J. J. Catal. 2013, 298 (0), 130–137.

(11) Kulkarni, N. V; Brennessel, W. W.; Jones, W. D. ACS Catal. 2017, 997–1002.

(12) Young, Z. D.; Hanspal, S.; Davis, R. J. ACS Catal. 2016, 6 (5), 3193–3202.

(13) Ho, C. R.; Shylesh, S.; Bell, A. T. ACS Catal. 2016, 6 (2), 939–948.

(14) Takahara, I.; Saito, M.; Inaba, M.; Murata, K. Catal. Letters 2005, 105 (3), 249–

252.

(15) Bi, J.; Guo, X.; Liu, M.; Wang, X. Catal. Today 2010, 149 (1), 143–147.

(16) Phung, T. K.; Busca, G. Catal. Commun. 2015, 68, 110–115.

(17) Zhang, X.; Wang, R.; Yang, X.; Zhang, F. Microporous Mesoporous Mater. 2008,

116 (1–3), 210–215.

(18) Han, Y.; Lu, C.; Xu, D.; Zhang, Y.; Hu, Y.; Huang, H. Appl. Catal. A Gen. 2011,

396 (1–2), 8–13.

(19) DeWilde, J. F.; Chiang, H.; Hickman, D. A.; Ho, C. R.; Bhan, A. ACS Catal. 2013,

3 (4), 798–807.

(20) Phung, T. K.; Proietti Hernández, L.; Lagazzo, A.; Busca, G. Appl. Catal. A Gen.

2015, 493, 77–89.

(21) Gayubo, A. G.; Tarrío, A. M.; Aguayo, A. T.; Olazar, M.; Bilbao, J. Ind. Eng.

Chem. Res. 2001, 40 (16), 3467–3474.

(22) Gayubo, A. G.; Alonso, A.; Valle, B.; Aguayo, A. T.; Olazar, M.; Bilbao, J. Fuel

2010, 89 (11), 3365–3372.

(23) Iwamoto, M. Molecules 2011, 16 (9), 7844–7863.

(24) Meng, T.; Mao, D.; Guo, Q.; Lu, G. Catal. Commun. 2012, 21 (0), 52–57.

(25) Song, Z.; Takahashi, A.; Nakamura, I.; Fujitani, T. Appl. Catal. A Gen. 2010, 384

(1–2), 201–205.

(26) Liu, C.; Sun, J.; Smith, C.; Wang, Y. Appl. Catal. A Gen. 2013, 467 (0), 91–97.

(27) Sun, J.; Zhu, K.; Gao, F.; Wang, C.; Liu, J.; Peden, C. H. F.; Wang, Y. J. Am.

Chem. Soc. 2011, 133 (29), 11096–11099.

Page 64: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

44

(28) Sun, J.; Baylon, R. A. L.; Liu, C.; Mei, D.; Martin, K. J.; Venkitasubramanian, P.;

Wang, Y. J. Am. Chem. Soc. 2016, 138 (2), 507–517.

(29) Lippits, M. J.; Nieuwenhuys, B. E. Catal. Today 2010, 154 (1), 127–132.

(30) Lippits, M. J.; Nieuwenhuys, B. E. J. Catal. 2010, 274 (2), 142–149.

(31) Tu, Y.-J.; Chen, Y.-W. Ind. Eng. Chem. Res. 1998, 37 (7), 2618–2622.

(32) Tu, Y.-J.; Li, C.; Chen, Y.-W. J. Chem. Technol. Biotechnol. 1994, 59 (2), 141–

147.

(33) Tu, Y.-J.; Chen, Y.-W.; Li, C. J. Mol. Catal. 1994, 89 (1), 179–189.

(34) Chang, F.-W.; Yang, H.-C.; Roselin, L. S.; Kuo, W.-Y. Appl. Catal. A Gen. 2006,

304 (0), 30–39.

(35) Chang, F.-W.; Kuo, W.-Y.; Lee, K.-C. Appl. Catal. A Gen. 2003, 246 (2), 253–

264.

(36) Gaspar, A. B.; Barbosa, F. G.; Letichevsky, S.; Appel, L. G. Appl. Catal. A Gen.

2010, 380 (1–2), 113–117.

(37) Inui, K.; Kurabayashi, T.; Sato, S. Appl. Catal. A Gen. 2002, 237 (1–2), 53–61.

(38) Inui, K.; Kurabayashi, T.; Sato, S.; Ichikawa, N. J. Mol. Catal. A Chem. 2004, 216

(1), 147–156.

(39) Sánchez, A. B.; Homs, N.; Fierro, J. L. G.; Piscina, P. R. de la. Catal. Today 2005,

107–108 (0), 431–435.

(40) Jørgensen, B.; Egholm Christiansen, S.; Dahl Thomsen, M. L.; Christensen, C. H.

J. Catal. 2007, 251 (2), 332–337.

(41) Santacesaria, E.; Carotenuto, G.; Tesser, R.; Di Serio, M. Chem. Eng. J. 2012, 179

(0), 209–220.

(42) Inui, T.; Matsuda, H.; Yamase, O.; Nagata, H.; Fukuda, K.; Ukawa, T.; Miyamoto,

A. J. Catal. 1986, 98 (2), 491–501.

(43) Palagin, D.; Sushkevich, V. L.; Ivanova, I. I. J. Phys. Chem. C 2016, 120 (41),

23566–23575.

(44) Chalk, S. G.; Miller, J. F. J. Power Sources 2006, 159 (1), 73–80.

(45) Taifan, W.; Baltrusaitis, J. Appl. Catal. B Environ. 2016, 198, 525–547.

(46) Ni, M.; Leung, M. K. H.; Leung, D. Y. C.; Sumathy, K. Renew. Sustain. Energy

Rev. 2007, 11 (3), 401–425.

(47) Matar, S.; Hatch, L. F. F. Chemistry of Petrochemical Processes; Elsevier, 2001.

(48) Yang, C.; Meng, Z. Y. J. Catal. 1993, 142 (1), 37–44.

(49) Kozlowski, J. T.; Davis, R. J. J. Energy Chem. 2013, 22 (1), 58–64.

(50) León, M.; Díaz, E.; Ordóñez, S. Catal. Today 2011, 164 (1), 436–442.

(51) Carvalho, D. L.; de Avillez, R. R.; Rodrigues, M. T.; Borges, L. E. P.; Appel, L. G.

Appl. Catal. A Gen. 2012, 415–416 (0), 96–100.

(52) Marcu, I.-C.; Tichit, D.; Fajula, F.; Tanchoux, N. Catal. Today 2009, 147 (3–4),

231–238.

(53) Tsuchida, T.; Kubo, J.; Yoshioka, T.; Sakuma, S.; Takeguchi, T.; Ueda, W. J.

Catal. 2008, 259 (2), 183–189.

(54) Quattlebaum, W. M.; Toussaint, W. J.; Dunn, J. T. J. Am. Chem. Soc. 1947, 69,

593–599.

(55) Toussaint, W. J.; Dunn, J. T.; Jachson, D. R. Ind. Eng. Chem. 1947, 39 (2), 120–

125.

(56) Lebedev, S. V. Zh Obs. Khim 1933, No. 3, 698–717.

Page 65: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

45

(57) Larina, O.; Kyriienko, P.; Soloviev, S. Catal. Letters 2015, 1–7.

(58) Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A.

ChemSusChem 2014, 7 (9), 2505–2515.

(59) Janssens, W.; Makshina, E. V. V.; Vanelderen, P.; De Clippel, F.; Houthoofd, K.;

Kerkhofs, S.; Martens, J. A. A.; Jacobs, P. A. A.; Sels, B. F. B. F. ChemSusChem

2014, 8 (6), 994–1008.

(60) Shylesh, S.; Gokhale, A. A.; Scown, C. D.; Kim, D.; Ho, C. R.; Bell, A. T.

ChemSusChem 2016, 9 (12), 1462–1472.

(61) Sushkevich, V. L.; Ivanova, I. I.; Taarning, E. Green Chem. 2015.

(62) Sushkevich, V. L.; Ivanova, I. I.; Ordomsky, V. V; Taarning, E. ChemSusChem

2014, 7 (9), 2527–2536.

(63) Taifan, W.; Yan, G. X.; Baltrusaitis, J. Catal. Sci. Technol. 2017, 7 (20), 4648–

4668.

(64) Ordomsky, V. V.; Sushkevich, V. L.; Ivanova, I. I. J. Mol. Catal. A Chem. 2010,

333 (1), 85–93.

(65) Phung, T. K.; Lagazzo, A.; Rivero Crespo, M. Á.; Sánchez Escribano, V.; Busca,

G. J. Catal. 2014, 311, 102–113.

(66) Taifan, W. E.; Bučko, T.; Baltrusaitis, J. J. Catal. 2017, 346, 78–91.

(67) Georgieff, K. K. J. Appl. Polym. Sci. 1966, 10 (9), 1305–1313.

(68) Wang, Z. In Comprehensive Organic Name Reactions and Reagents; John Wiley

& Sons, Inc., 2010.

(69) Ostromislenskiy, I. J. Russ. Phys. Chem. Soc. 1915, No. 47, 1472 –1506.

(70) Gruver, V.; Sun, A.; Fripiat, J. J. Catal. Letters 1995, 34 (3–4), 359–364.

(71) Chieregato, A.; Velasquez Ochoa, J.; Bandinelli, C.; Fornasari, G.; Cavani, F.;

Mella, M. ChemSusChem 2014, 8 (2), 377–388.

(72) Makshina, E. V; Dusselier, M.; Janssens, W.; Degreve, J.; Jacobs, P. A.; Sels, B. F.

Chem. Soc. Rev. 2014, 43 (22), 7917–7953.

(73) Pomalaza, G.; Capron, M.; Ordomsky, V.; Dumeignil, F. Catalysts 2016, 6 (12),

203–237.

(74) Makshina, E. V.; Janssens, W.; Sels, B. F.; Jacobs, P. A. Catal. Today 2012, 198

(1), 338–344.

(75) Niiyama, H.; Morii, S.; Echigoya, E. Bull. Chem. Soc. Jpn. 1972, 45 (3), 655–659.

(76) Chung, S.-H.; Angelici, C.; Hinterding, S. O. M.; Weingarth, M.; Baldus, M.;

Houben, K.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ACS Catal. 2016, 6 (6),

4034–4045.

(77) Huang, X.; Men, Y.; Wang, J.; An, W.; Wang, Y. Catal. Sci. Technol. 2017, 7 (1),

168–180.

(78) Jones, M. D.; Keir, C. G.; Iulio, C. Di; Robertson, R. A. M.; Williams, C. V;

Apperley, D. C. Catal. Sci. Technol. 2011, 1 (2), 267–272.

(79) De Baerdemaeker, T.; Feyen, M.; Müller, U.; Yilmaz, B.; Xiao, F.-S.; Zhang, W.;

Yokoi, T.; Bao, X.; Gies, H.; De Vos, D. E. ACS Catal. 2015, 3393–3397.

(80) Bhattacharyya, S. K.; Sanyal, S. K. J. Catal. 1967, 7 (2), 152–158.

(81) Corson, B. B.; Jones, H. E.; Welling, C. E.; Hinckley, J. A.; Stahly, E. E. Ind. Eng.

Chem. 1950, 42 (2), 359–373.

(82) Muller, P.; Burt, S. P.; Love, A. M.; McDermott, W. P.; Wolf, P.; Hermans, I. ACS

Catal. 2016, 6 (10), 6823–6832.

Page 66: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

46

(83) Jones, M. D. Chem. Cent. J. 2014, 8 (1), 53.

(84) Posada, J. A.; Patel, A. D.; Roes, A.; Blok, K.; Faaij, A. P. C.; Patel, M. K.

Bioresour. Technol. 2013, 135, 490–499.

(85) Bhattacharyya, S. K.; Avasthi, B. N. Ind. Eng. Chem. Process Des. Dev. 1963, 2

(1), 45–51.

(86) Hayashi, Y.; Akiyama, S.; Miyaji, A.; Sekiguchi, Y.; Sakamoto, Y.; Shiga, A.;

Koyama, T.; Motokura, K.; Baba, T. Phys. Chem. Chem. Phys. 2016, 18 (36),

25191–25209.

(87) Velasquez Ochoa, J.; Bandinelli, C.; Vozniuk, O.; Chieregato, A.; Malmusi, A.;

Recchi, C.; Cavani, F. Green Chem. 2016, 18, 1653–1663.

(88) Velasquez Ochoa, J.; Malmusi, A.; Recchi, C.; Cavani, F.; Ochoa, J. V.; Malmusi,

A.; Recchi, C.; Cavani, F. ChemCatChem 2017, 9 (12), 2128–2135.

(89) Larina, O. V.; Kyriienko, P. I.; Trachevskii, V. V.; Vlasenko, N. V.; Soloviev, S.

O. Theor. Exp. Chem. 2016, 51 (6), 387–393.

(90) Ohnishi, R.; Akimoto, T.; Tanabe, K. J. Chem. Soc. Chem. Commun. 1985, No.

22, 1613–1614.

(91) Da Ros, S.; Jones, M. D.; Mattia, D.; Pinto, J. C.; Schwaab, M.; Noronha, F. B.;

Kondrat, S. A.; Clarke, T. C.; Taylor, S. H. ChemCatChem 2016, 8, 1–12.

(92) Sushkevich, V. L.; Ivanova, I. I. ChemSusChem 2016, 9 (16), 2216–2225.

(93) Baylon, R. a. L.; Sun, J.; Wang, Y. Catal. Today 2015, 1–7.

(94) Larina, O. V; Kyriienko, P. I.; Soloviev, S. O. Theor. Exp. Chem. 2016, 52 (1), 51–

56.

(95) Kyriienko, P. I.; Larina, O. V; Soloviev, S. O.; Orlyk, S. M.; Calers, C.; Dzwigaj,

S.; Pisarzhevsky, L. V. ACS Sustain. Chem. Eng. 2017, 5 (3), 2075–2083.

(96) Natta, G.; Rigamonti, R. Chim. E Ind. 1947, 29, 195–201.

(97) Ruckenstein, E.; Khan, A. Z. J. Catal. 1993, 141 (2), 628–647.

(98) Cavalleri, M.; Pelmenschikov, A.; Morosi, G.; Gamba, A.; Coluccia, S.; Martra, G.

In Oxide-based Systems at the Crossroads of ChemistrySecond International

Workshop October 8-11, 2000, Como, Italy; A. Gamba, C. C. and S. C. B. T.-S. in

S. S. and C., Ed.; Elsevier, 2001; Vol. Volume 140, pp 131–139.

(99) Kvisle, S.; Aguero, A.; Sneeden, R. P. A. Appl. Catal. 1988, 43 (1), 117–131.

(100) Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A. Catal.

Sci. Technol. 2015.

(101) Natta, G.; Rigamonti, R. Chim. Ind. 1947, 29, 239–243.

(102) Lewandowski, M.; Babu, G. S.; Vezzoli, M.; Jones, M. D.; Owen, R. E.; Mattia,

D.; Plucinski, P.; Mikolajska, E.; Ochenduszko, A.; Apperley, D. C. Catal.

Commun. 2014, 49, 25–28.

(103) Kitayama, Y.; Michishita, A. J. Chem. Soc., Chem. Commun. 1981, No. 9, 401–

402.

(104) Kitayama, Y.; Satoh, M.; Kodama, T. Catal. Letters 1996, 36 (1–2), 95–97.

(105) Sekiguchi, Y.; Akiyama, S.; Urakawa, W.; Koyama, T.; Miyaji, A.; Motokura, K.;

Baba, T. Catal. Commun. 2015, 68, 20–24.

(106) Angelici, C.; Meirer, F.; van der Eerden, A. M. J.; Schaink, H. L.; Goryachev, A.;

Hofmann, J. P.; Hensen, E. J. M.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ACS

Catal. 2015, 5 (10), 6005–6015.

(107) Sushkevich, V. L.; Palagin, D.; Ivanova, I. I. ACS Catal. 2015, 4833–4836.

Page 67: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

47

(108) Sushkevich, V. L.; Ivanova, I. I. Appl. Catal. B Environ. 2017, 215, 36–49.

(109) Sushkevich, V. L.; Ivanova, I. I.; Taarning, E. ChemCatChem 2013, 5 (8), 2367–

2373.

(110) Patil, P. T.; Liu, D.; Liu, Y.; Chang, J.; Borgna, A. Appl. Catal. A Gen. 2017, 543,

67–74.

(111) Bhattacharyya, S. K.; Ganguly, N. D. J. Appl. Chem. 1962, 12 (3), 97–104.

(112) Kim, T.-W.; Kim, J.-W.; Kim, S.-Y.; Chae, H.-J.; Kim, J.-R.; Jeong, S.-Y.; Kim,

C.-U. Chem. Eng. J. 2014.

(113) Chae, H.-J.; Kim, T.-W.; Moon, Y.-K.; Kim, H.-K.; Jeong, K.-E.; Kim, C.-U.;

Jeong, S.-Y. Appl. Catal. B Environ. 2014, 150–151, 596–604.

(114) Müller, P.; Wang, S.-C.; Burt, S. P.; Hermans, I. ChemCatChem 2017, 9 (18),

3572–3582.

(115) Branda, M. M.; Rodríguez, A. H.; Belelli, P. G.; Castellani, N. J. Surf. Sci. 2009,

603 (8), 1093–1098.

(116) Jensen, F. Introduction to Computational Chemistry, 2nd ed.; John Wiley & Sons

Ltd., 2006.

(117) Weckhuysen, B. M. Phys. Chem. Chem. Phys. 2003, 5 (20), 4351–4360.

(118) Chakrabarti, A.; Ford, M. E.; Gregory, D.; Hu, R.; Keturakis, C. J.; Lwin, S.;

Tang, Y.; Yang, Z.; Zhu, M.; Bañares, M. A.; Wachs, I. E. Catal. Today 2017,

283, 27–53.

(119) Bañares, M. A. Catal. Today 2005, 100 (1), 71–77.

(120) Jentoft, F. C. B. T.-A. in C. Academic Press, 2009; Vol. 52, pp 129–211.

(121) Schoonheydt, R. A. Chem. Soc. Rev. 2010, 39 (12), 5051–5066.

(122) Bravo-Suarez, J. J.; Subramaniam, B.; Chaudhari, R. V. J. Phys. Chem. C 2012,

116 (34), 18207–18221.

(123) Gorte, R. J. Catal. Today 1996, 28 (4), 405–414.

(124) Redhead, P. a. Vacuum 1962, 12 (4), 203–211.

(125) ter Veen, H. R. J.; Kim, T.; Wachs, I. E.; Brongersma, H. H. Catal. Today 2009,

140 (3), 197–201.

(126) McWilliam, I. G.; Dewar, R. A. Nature 1958, 181, 760.

(127) Doerr, A. Nat. Methods 2008, 6, 34.

Page 68: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

48

Chapter 2

Experimental Methods

1. Introduction ........................................................................................................ 48

2. Computational Details ....................................................................................... 49

2.1. Electronic structure calculations .............................................................. 49

2.2. Structural optimization calculations ......................................................... 49

2.3. Structural model ....................................................................................... 50

2.4. Free-energy calculation ............................................................................ 52

2.4.1. Working equations ................................................................................... 52

2.4.2. Partitioning of atomic degrees of freedom in interacting and non-

interacting systems ............................................................................................... 53

2.4.3. Calculation of harmonic vibrational frequencies ..................................... 54

3. Experimental Methods ...................................................................................... 55

3.1. Catalyst synthesis ....................................................................................... 55

3.1.1. Synthesis of magnesium oxide, MgO, catalyst ......................................... 55 3.1.2. Synthesis of MgO/SiO2 catalysts .............................................................. 55

3.1.3. Synthesis of promoted wet-kneaded MgO/SiO2 catalysts ......................... 56

3.2. Catalytic reactivity study ........................................................................... 56

3.3. Catalyst characterization ............................................................................ 58

3.3.1. High-sensitivity low energy ion scattering (HS-LEIS) .............................. 58

3.3.2. XRD and BET surface area ........................................................................59

3.3.3. Transition metal concentration measurements ........................................... 59

3.3.4. Scanning transmission electron microscopy .............................................. 60

3.3.5. In-situ spectroscopy ................................................................................... 60

3.3.6. Acid-base characterization using pyridine, NH3, CO2, and methanol as

probe molecules .................................................................................................... 61

3.4. Reaction mechanism study using in-situ DRIFTS spectroscopy and TPRS

............................................................................................................................... 62

3.5. Operando XANES and EXAFS spectroscopy during ethanol reaction to

1,3-BD over Cu- and Zn-promoted MgO/SiO2 catalysts ...................................... 64

References ....................................................................................................................... 65

1. Introduction

This chapter is dedicated for both computational details and experimental methods

used throughout the research work.

Page 69: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

49

2. Computational Details

2.1. Electronic structure calculations.

Periodic DFT calculations have been performed using the VASP code.1–4 The

Kohn–Sham equations have been solved variationally in a plane-wave basis set using the

projector-augmented-wave (PAW) method of Blochl,5 as adapted by Kresse and Joubert.5

The exchange-correlation energy was described by the PBE generalized gradient

approximation.6 Brillouin-zone was sampled using 2x2x1 k-point mesh. The plane-wave

cutoff was set to 400 eV. The convergence criterion for the electronic self-consistency

cycle, measured by the change in the total energy between successive iterations, was set to

10-6 eV/cell.

2.2. Structural optimization calculations.

Transition states have been identified using the DIMER method,7 as improved by

Heyden et al.8 Atomic positions were considered to be relaxed if all forces acting on the

atoms were smaller than 0.005 eV/Å. Transition states were proven to be first-order saddle

points of the potential energy surface using vibrational analysis. The intrinsic reaction

coordinates9,10 (IRCs) for the forward and backward reaction steps were identified using

the damped velocity Verlet algorithm.11 The structures corresponding to potential energy

minima along the IRC were further relaxed using a conjugate-gradient algorithm such as

to satisfy the same optimization criterion as for transition states. Vibrational analysis was

performed to ensure that the relaxed structures correspond to true potential energy minima.

This procedure guarantees that reactant and product states are linked by a path with a single

transition state.

Page 70: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

50

The Gibbs free-energy calculations have been performed using the harmonic/rigid

rotor approximation to the transition state theory12 for the temperature of 723 K. The

detailed calculations for the Gibbs energy estimations are provided in subchapter 2.4. All

relative energies were referenced to the sum of the relaxed MgO slab, and three molecules

of gas phase ethanol. The first-order reaction constants were computed using:

𝑘(𝑇) =𝑘𝐵𝑇

ℎ𝑒−

∆𝐺𝐴(𝑇)

𝑅𝑇 (2.1)

where kB – 1.380662x10-23 m2 kg/s2 K1, T = 723 K, h - 6.626176x10-34 m2 kg/s, R - 1.987

cal/K mol,and GA(T) is molar Gibbs free-energy of activation defined for the forward

and reverse reactions as follows:

𝛥𝐺𝐴,𝑓𝑜𝑟𝑤𝑎𝑟𝑑 = 𝐺𝑇𝑆 − 𝐺𝐼𝑆 (2.2)

𝛥𝐺𝐴,𝑟𝑒𝑣𝑒𝑟𝑠𝑒 = 𝐺𝑇𝑆 − 𝐺𝐹𝑆 (2.3)

with subscripts TS, FS and IS representing transition state, final state, and initial state,

respectively. Similarly, the reaction free energy (𝛥𝐺𝑅𝑥) is defined as follows:

𝛥𝐺𝑅𝑥 = 𝐺𝐹𝑆 − 𝐺𝐼𝑆 (2.4)

The rate constants for the forward and reverse reaction steps are labelled as kf and kr,

respectively.

2.3. Structural model.

Under normal conditions, MgO crystallizes in the rocksalt crystal structure, with

(001) being the most prominent surface. For the specific reaction of ethanol to 1,3-

butadiene, the original working catalyst is MgO/SiO2 13,14. The addition of MgO to SiO2,

regardless of the methods used to synthesize the catalysts, does not result in the formation

of MgSiO4 solid solution.13,15–17 It was found that MgO crystalline phase was the only

phase found on the amorphous surface. In contrast, a study by Angelici et al.18 claims that

Page 71: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

51

MgSiO4 phase was indeed formed and it was suggested to be responsible for the aldol

condensation of the acetaldehyde intermediates. The authors argued that FTIR shows the

presence of a hydroxyl group which coincides with that found in talc. This statement

contradicts their previous study, in which the presence of the silicate was not confirmed by

XRD.19 The addition of MgO to amorphous silica combined with increased surface area of

the catalyst is also known to create more defects on the MgO itself, which clearly

contributes to the higher activity and selectivity of the catalyst. The unsaturated surface

oxygen atoms, e.g. corner and edge O, are unfavorable energetically leading to the

tendency to release an electron and to turn into O- species. Hence these unsaturated oxygen

atoms act as donors of an electron pair, i.e. as Lewis bases.20 Cube corners,21 terraces, steps,

corners, and reverse corners were studied extensively.22

Periodic three layer slab of MgO consisting of 8 x 6 x 3 primitive cells was used

throughout all the calculations. Previous investigations have identified defect sites of MgO

catalysts, most notably the three- and four-fold coordinated surface atoms (Mg2+3C O2-

4C),

Figure 2.1 Periodic MgO slab used throughout the calculations. The whited out bottom

layer indicates the atoms whose positions were kept frozen during calculations.

Page 72: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

52

as the active sites22–27 whereas atoms at surface terraces were found to be relatively

unreactive.28 The positions of atoms in the bottommost layer were fixed while remaining

atoms were relaxed. PBE lattice parameter of 4.255 Å was used to construct the slab. This

value is similar to values of 4.261 (PBE) and 4.186 Å (experimental).29 In this work we

considered the Mg3C coordinated to O4C in the form of stepped kink, depicted in Figure 2.1,

as the active sites for the reaction. The large slab is used to accommodate several

intermolecular reactions that might include interactions of C2 and C4 intermediates and

also to avoid significant interaction between molecules in the neighboring unit cells.

2.4. Free-energy calculation

The free-energy calculations have been performed using a static approach based on

harmonic and rigid rotor approximations to vibrational and rotational degrees of freedom.

Although this methodology is described in detail in many textbooks,12 we find it useful to

summarize the working equations in this text.

2.4.1. Working equations

Within the static approximation used in this study, Gibbs free energy for each state

is expressed as a sum of contribution of electronic (el), vibrational (vib), rotational (rot),

and translational (tr) degrees of freedom (DOF):

𝐺 = 𝐺𝑒𝑙 + 𝐺𝑣𝑖𝑏 + 𝐺𝑟𝑜𝑡 + 𝐺𝑡𝑟 (2.5)

Discussion of partitioning of atomic degrees of freedom in the case of interacting and non-

interacting systems is provided in section 1.2.

The electronic contribution for a system in a singlet electronic state (all systems discussed

in this work) takes a form:

𝐺𝑒𝑙 = 𝐸𝐷𝐹𝑇 (2.6)

Page 73: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

53

where 𝐸𝐷𝐹𝑇 is a Kohn-Sham energy computed by solving Schrödinger equation at the DFT

level. The vibrational contribution is expressed as

𝐺𝑣𝑖𝑏 =3

2𝑘𝐵𝑇 − 𝑇𝑘𝐵 ∑ (

ℎ𝜈𝑖

𝑘𝐵𝑇

1

𝑒

ℎ𝜈𝑖𝑘𝐵𝑇−1

− ln (1 − 𝑒−

ℎ𝜈𝑖𝑘𝐵𝑇))

𝑁𝑣𝑖𝑏𝑖=1 (2.7)

where 𝑘𝐵 and ℎ are fundamental constants (Boltzmann and Planck, respectively), 𝑇 is

thermodynamic temperature (723 K), 𝑁𝑣𝑖𝑏 is the number of vibrational degrees of freedom,

and 𝜈 is a harmonic vibrational frequency. The rotational contribution is expressed as

𝐺𝑟𝑜𝑡 = −𝑘𝐵𝑇 ln (√𝜋

𝜎(

8𝜋2𝑘𝐵𝑇

ℎ2 )3/2

√𝐼1𝐼2𝐼3) (2.8)

where 𝜎 is a symmetry index, and 𝐼1, 𝐼2, and 𝐼3 are moments of inertia of a molecule. The

values of 𝜎 for molecules considered in this study are compiled in Tab.S1. Finally, the term

𝐺𝑡𝑟 is expressed as follows

𝐺𝑡𝑟 = −𝑘𝐵𝑇 ln (𝑉𝑚 (2𝜋𝑀𝑘𝐵𝑇

ℎ2 )3/2

) (2.9)

where 𝑉𝑚 = 𝑘𝐵𝑇/𝑝 is a volume occupied by one particle of ideal gas at given external

pressure 𝑝 (101325 Pa) and temperature (723 K), and 𝑀 is the total molecular mass.

Table 2.1. Symmetry indices 𝜎 for gas-phase molecules considered in this study.

Molecule 𝝈

acetaldehyde 1

crotonaldehyde 1

ethanol 1

butadiene 2

dihydrogen 2

water 2

2.4.2. Partitioning of atomic degrees of freedom in interacting and non-interacting

systems

For a system consisting of 𝑁𝑠,𝑓𝑟𝑒𝑒 free substrate atoms (in our model 𝑁𝑠,𝑓𝑟𝑒𝑒

corresponds to all substrate atoms except of the bottommost layer of the slab, see

Page 74: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

54

subchapter 2.3 and Figure 2.1) and 𝑁𝑀 atoms forming molecules adsorbed on the substrate,

the number of vibrational degrees of freedom is 𝑁𝑣𝑖𝑏 = 3(𝑁𝑠,𝑓𝑟𝑒𝑒 + 𝑁𝑀). There are no

rotational and translational degrees of freedom in such a case and hence also the

contribution of 𝐺𝑟𝑜𝑡 and 𝐺𝑡𝑟 to Gibbs free energy is zero. In the case of systems consisting

of a substrate and 𝑃 molecules that neither interact with each other nor they interact with

the substrate, the total number of vibrational degrees of freedom is 3𝑁𝑠,𝑓𝑟𝑒𝑒 +

∑ (3𝑁𝑀𝑖− 3 − 𝑁𝑟𝑜𝑡,𝑖)

𝑃𝑖=1 , where 𝑁𝑀𝑖

and 𝑁𝑟𝑜𝑡,𝑖 are the number of atoms and the number

of rotational degrees of freedom in the molecule i. Each molecule has 3 translational and 2

(linear molecules) or 3 (nonlinear molecules) rotational degrees of freedom.

2.4.3. Calculation of harmonic vibrational frequencies

The harmonic vibrational frequencies have been computed using the finite

differences method implemented in VASP. The numerical differentiation has been done

using the a differences formula with displacement of size 0.02 Å. Even the use of quite

stringent relaxation criterion (maximal force smaller than 0.005 eV/ Å) does not ensure the

correct eigenvalue spectrum of dynamical matrix (i.e. zero imaginary vibrational

frequencies in the case of minima and one imaginary vibrational frequency in the case of

first-order saddle) in all cases. In order to obtain correct vibrational spectrum in such a

problematic case, several iterations consisting of line-minimization of energy along the

incorrect unstable directions, followed by a full relaxation of the atomic positions and

dynamical matrix calculations was performed.

Page 75: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

55

3. Experimental Methods

3.1. Catalyst synthesis

3.1.4. Synthesis of magnesium oxide, MgO, catalyst

MgO catalyst was synthesized using a modified thermal decomposition method.30

In a typical synthesis, 2.23g (8.7 mmol) of Mg(NO3)26H2O (Sigma-Aldrich) were

dissolved in 40 ml methanol and then 1 ml water was added. A 30 ml methanol solution

containing 0.7 g (17.4 mmol) of NaOH was added drop-wise to the resulting solution under

reflux temperature. After 30 minutes a white precipitate was collected by centrifugation

(Thermo Sorvall™ Legend™ XT). The isolated precipitate was washed three times using a

1:1 ratio solution of ethanol/water and then separated using centrifugation. The resulting

wet samples were dried at 80 °C overnight. The resulting dry magnesium hydroxide solid

was ground using a mortar and pestle and calcined at 800 °C in a calcination oven (Thermo

Lindberg™ Blue M). Here, a ramping rate of 10 °C/min for 3 hours was used under an

oxidizing atmosphere with an air flow rate of 50 ml/min. Natural convection was used to

cool down the samples.

3.1.5. Synthesis of MgO/SiO2 catalysts

Two methods of preparation are investigated in this study, i.e. incipient wetness

impregnation (IWI) and wet-kneaded (WK). The incipient wetness impregnation was done

using final Mg/Si mass ratio of 1. Precursor used in this method is Mg(NO3)26H2O (Sigma-

Aldrich) in water, impregnated on fumed silica (Cabot). The solid is then dried overnight

under ambient condition, followed by drying at 80°C overnight, before further calcined at

800 °C. The wet-kneaded MgO/SiO2 catalysts were prepared by utilizing some of the

magnesium hydroxide material obtained in Section 3.1.1 by a thermal decomposition

Page 76: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

56

method before its calcination. Instead of calcining, the hydroxide was wet-kneaded with

fumed silica (Cabot).13 The corresponding amounts of silica and magnesium hydroxide

were wet-kneaded in deionized water for 4 hours, centrifuged, dried overnight at 80 °C and

calcined. In Chapter 4 and 5, 800°C was chosen as calcination temperature for both

catalysts. In Chapter 4, the catalyst is labeled as WK (1:1), while in Chapter 5, the catalysts

are labeled as MgSi-WK and MgSi-IWI for comparing between WK and IWI methods,

and a WK catalyst calcined at 500°C is labeled as MgSi-WK2. In Chapter 6, for comparison

between unpromoted and promoted WK catalysts calcined at 500°C, the unpromoted

catalyst is simply referred to as MgSi.

3.1.6. Synthesis of promoted wet-kneaded MgO/SiO2 catalysts.

Following synthesis the wet-kneaded MgO/SiO2 (1:1) catalyst, drying is instead

carried out at room temperature overnight, and the catalyst was then impregnated with

transition metals, i.e., Cu or Zn. Copper nitrate trihydrate (Alfa Aesar) and zinc nitrate

hexahydrate (Sigma) were used as precursors. The catalysts were then dried at room

temperature overnight and further calcined at 500 °C for 3 hours. For comparison purpose,

an unpromoted catalyst is also calcined at 500 °C for 3 hours. These catalysts are labeled

as CuMgSi and ZnMgSi, while the corresponding binary reference catalysts, e.g. Cu(Zn)-

SiO2, Cu(Zn)-MgO, are labeled as CuSi, ZnSi, CuMg and ZnMg, respectively.

3.2. Catalytic reactivity study

The catalytic tests were performed in a Microactivity-Reference fixed bed reactor

from PID Eng Tech (Spain). A quartz tube reactor was used with the quartz wool positioned

so as to support the catalyst bed (0.1 g, pelletized, crushed and sieved to 100-150 µm

particle size). Additional SiO2 powder (Sigma) was used to increase the bed length so as

Page 77: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

57

to maintain the plug flow condition. Ethanol was delivered with helium gas by bubbling

the gas through a chilled ethanol saturator with a total flow of 50 ml/min. The bubbler

temperature was varied to manipulate the weight hourly space velocity (WHSV). The hot

box temperature in the reactor was set at 100 °C to prevent any reactant or product

condensation. Prior to the reaction, the catalyst was activated by heating it to 500 °C at a

heating rate of 10 °C/min and then held at that temperature for 1 hour under 30 ml/min He

flow. The reactions were performed at 375 °C. The products were kept in the vapor phase

and then analyzed using a gas chromatograph equipped with an FID detector and Restek

RT-Q-Bond column. The reactant ethanol and principal products, i.e., ethylene,

acetaldehyde and 1,3-BD, were quantified based on the calibration carried out using a

standard reference mixture (Praxair).

Titration experiment was carried out to poison both basic and acidic sites. To poison

basic sites, probe molecules, i.e. CO2 and propionic acid, were used. Poisoning acidic sites

were carried out by using NH3 as the probe molecule, and by post-treatment using NaOH.

For this post-treatment method, the catalyst was impregnated with a very dilute NaOH

solution, with final catalysts containing 250, 500, and 1000 ppm NaO, and let dry at room

temperature without further thermal treatment. In a typical titration experiment, the catalyst

was let to achieve a steady-state condition at a selected WHSV and reaction temperature.

Probe molecules were then co-fed into the reactor using MFC for CO2 and 1% NH3 in N2,

while propionic acid was delivered using a chilled saturator containing mixture of

propionic acid/ethanol (3:7). After a new steady-state is achieved, the feed was reverted

back to only ethanol to check for the activity recovery.

Page 78: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

58

3.3. Catalyst characterization

Unpromoted catalysts, i.e. MgSi-IWI and MgSi-WK, were characterized using

HS-LEIS, XRD, BET surface area measurement, and combination of in-situ IR and UV-

Vis measurements. TPRS experiments were also run using MgSi-WK. Transition metal-

promoted catalysts were characterized using XRD and BET surface area measurement,

ICP-OES, XPS, STEM, in-situ IR and UV-Vis, and also operando XAS experiments.

3.3.1. High-sensitivity low energy ion scattering (HS-LEIS)

The unpromoted IWI and WK catalysts (1:1), calcined at 800 °C, were prepared for

analysis by dispersing into an appropriate sample crucible for a heatable sample holder for

the LEIS spectrometer, ION-TOF Qtac100, and then compacting it with a sample press. The

crucible was then affixed to a sample holder with an integrated cartridge heater and a

thermocouple was placed in a hole on the crucible.

After being placed in vacuum, the temperature of a sample was raised to about 50°C

for outgassing. O2 was then introduced into the preparation temperature at an unmeasured

pressure likely between 100 and 300 mbar. The temperature of the sample was then

increased at a rate of 10°C/min to a maximum temperature of 500°C. This temperature was

held for 60 min, at which time the temperature was allowed to decrease and the preparation

chamber was evacuated. The sample was then transferred into the analysis chamber.

Charge neutralization was invoked during spectra acquisition and sputtering. For primary

ion beam, the following parameter was used: 3.0 keV He+, 1500 1500 m raster, at 2 x

1014 ions cm-2 cyc-1, 3000 eV pass energy. The following conditions were applied during

sputtering: 1.0 keV Ar+, 2000 x 2000 m raster, 5 x 1014 ions cm-2 cyc-1.

Page 79: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

59

3.3.2. XRD and BET surface area

Bulk structural information of the catalysts was characterized using XRD. XRD

patterns were obtained using PANalytical Empyrean powder X-ray diffractometer using

Cu Kα1,2 with λ=1.5418 Å operating at 45 kV. Measurements were carried out between

2θ=10° and 100° using a step size of 0.05°. The BET specific surface areas of the catalysts

were determined by nitrogen adsorption at 77 K on a Micromeritics ASAP 2010

instrument. All samples were degassed under nitrogen flow at 623 K for 12 h before the

measurements.

3.3.3. Transition metal concentration measurements.

The weight transition metal concentration of Cu- and Zn-promoted MgO/SiO2

catalysts was determined using Inductively Coupled Plasma-Optical Emission

Spectroscopy (ICP-OES, PerkinElmer Optima 2000 DV). About 10 mg of catalyst was

digested using 40 ml solution containing 1:1:1 H2O, HCl and HNO3. Cu concentration was

measured to be 0.8%, similar to that used by Weckhuysen and coworkers18,19 while Zn was

2.5%, close to that reported by Larina et al.15

The XPS measurements were carried out to corroborate the results of ICP-OES with

a PHI 5600ci instrument using a non-monochromatized Al Kα X-ray source. The pass

energy of the analyzer was 58.7 eV, acquisition area with diameter of ~800 um and the

scan step size was 0.125 eV. Binding energies were corrected for charging by referencing

to the C 1s peak at 284.8 eV. Atomic concentrations were calculated from the areas under

individual high-resolution XPS spectra using manufacturer-provided sensitivity factors.

Page 80: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

60

3.3.4. Scanning transmission electron microscopy

The morphology of the catalyst particles was investigated using a dedicated

Scanning Transmission Electron Microscope (STEM) (Hitachi 2700C) operating at 200

kV.

3.3.5. In-situ spectroscopy

Diffuse Reflectance Infrared Spectroscopy (DRIFTS) was used to probe the

composition and changes in hydroxyl (OH) groups on the catalyst surface under dehydrated

conditions. A Thermo Nicolet iS50 infrared spectrometer equipped with a Mercury-

Cadmium-Tellurium (MCT) liquid nitrogen cooled detector was used in combination with

a Harrick Praying Mantis™ diffuse reflection accessory equipped with ZnSe windows.

About 30 mg of <100 µm catalyst samples were loaded into the DRIFTS cell. The smaller

particle size was used to ensure a uniform catalyst bed surface for spectroscopy. Similar to

the steady state reaction testing, the catalyst activation was carried out by heating it up to

500 °C at 10 °C/min and keeping it at that temperature for 1 hour under 30 ml/min He flow.

The catalyst was then cooled down to 100 °C under 30 ml/min N2 (Praxair) flow. During

the cooling reference spectra of the catalysts were acquired at 400°C, 300°C, 200°C and

100 °C. All spectra were averaged over 96 scans at a resolution of 4 cm-1.

In-situ UV-vis DRS measurements were performed using an Agilent Technologies

Cary 5000 UV-Vis- NIR spectrophotometer equipped with a Praying Mantis TM diffuse

reflection accessory. Finely ground samples (< 100µm) of supported catalyst powders were

loaded into the environmental cell (Harrick, HVC-DR2) and then UV-vis spectra were

collected in the 200-800 nm region. An MgO reflectance standard was used as the baseline.

The experimental protocol used for DRIFTS experiments was also used in the in-situ UV-

Page 81: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

61

VIS DRS experiments. The Kubelka-Munk function was calculated from the absorbance

of the UV-vis DRS. The edge energy (Eg) for allowed transitions was determined by

finding the intercept between the straight line and the abscissa on the Tauc plot derived

from the UV-Vis spectra.

3.3.6. Acid-base characterization using pyridine, NH3, CO2, and methanol as probe

molecules

A Thermo Nicolet iS50 infrared spectrometer equipped with a Mercury-Cadmium-

Tellurium (MCT) liquid nitrogen cooled detector was used with a Harrick Praying

Mantis™ diffuse reflection accessory and ZnSe windows to study the acidity and

basicity of the catalyst. About 30 mg of sample was pressed and loaded into the DRIFTS

cell. Catalyst activation was carried out by heating it up to 773 K at a rate of 10 K/min and

then held at that temperature for 1 hour under 30 ml/min air flow, in agreement with the

fixed bed experiment catalyst preparation procedure. The catalyst was then cooled down

to 373 K under 30 ml/min nitrogen (Praxair) flow. During the cooling, reference spectra of

the catalysts were acquired every 50 K. All spectra were averaged over 96 scans at a

resolution of 4 cm-1. Probe molecules, i.e. NH3, CO2 and pyridine, were used to characterize

the acidity and basicity of the catalyst. In general, the probe molecule is adsorbed on the

surface for 15 minutes shortly after the catalyst temperature is brought down to 373 K. This

step is followed by extensive purging using 30 ml/min N2 (Praxair) for 45 minutes. Spectra

were then continuously recorded every minute during which time the temperature was

increased to 723 K under 30 ml/min N2 flow. CO2 and NH3 (Praxair) gas cylinder is used

for delivery method, while pyridine delivery method involved bubbling N2 through the

pyridine saturator.

Page 82: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

62

Methanol operando temperature programmed DRIFTS-MS sample preparation

was carried out in a similar manner. The product was continuously monitored using a

Cirrus 2 benchtop atmospheric pressure gas analysis system (MKS Instruments). Methanol

was used because it can test and yield products formed at the acidic, basic and redox sites.31

Briefly, after the catalyst activation step, the CH3OH was preadsorbed on the sample

surface as a saturated vapor at 4 °C using 50 ml/min helium as a carrier gas with a cell

temperature of 100 °C for 30 minutes. The catalyst was subsequently flushed with pure

helium at 30 ml/min for 1 hour. Spectra were then continuously recorded every minute,

while the temperature was ramped up to 450 °C at a rate of 10 °C/min under He flow.

Unless stated otherwise, reference spectra obtained at the corresponding temperatures were

subtracted from the acquired spectra to eliminate contribution from the catalysts.

Calibration of methanol and CO2 was performed using a mixture of both products with He

at different concentrations, while formaldehyde - by reactive calibration of methanol

dehydrogenation over Cu/SiO2 catalyst. The reaction was kept at low conversion to limit

the occurrence of secondary reactions, forming such molecules as dimethoxymethane and

methyl formate. A mass balance for the reaction system was then calculated to determine

the response factor of the formaldehyde.

3.4. Reaction mechanism study using in-situ DRIFTS spectroscopy and TPRS

A Thermo Nicolet iS50 infrared spectrometer equipped with a Mercury-Cadmium-

Tellurium (MCT) liquid nitrogen cooled detector was used with a Harrick Praying

Mantis™ diffuse reflection accessory and ZnSe windows to study the nature of the

hydroxyl groups, as well as the adsorbates on the catalyst surface. Catalyst activation is

carried out according to the procedure mentioned in Section 2.3. To monitor reactive

Page 83: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

63

surface intermediates, ethanol was pre-adsorbed onto the sample surface as a saturated

vapor at 298 K using 30 ml/min nitrogen as a carrier gas at 373 K for 20 minutes.

Physisorbed molecules were removed with pure nitrogen at 30 ml/min for 40 minutes.

Liquid ethanol (200 proof, Koptec), crotonaldehyde (Acros organics, +99%) and crotyl

alcohol were used (Sigma, 96%). For acetaldehyde DRIFTS experiments, a gaseous

mixture of 5% acetaldehyde in nitrogen (Praxair) was used. Crotonaldehyde and crotyl

alcohol were handled with extra caution due to their toxicity. In particular, transporting the

chemical was done in the hood to an enclosed, chilled bubbler (2-4°C). The enclosed,

chilled bubbler was then installed to the gas flow delivery system while still being chilled.

Chilled bubbler lowered the partial pressure of crotonaldehyde and crotyl alcohol, further

limiting exposure to the vapor. Spectra were then continuously recorded every minute

during which time the temperature was increased to 723 K with or without the continuous

vapor flow of the reactants.

Four types of infrared spectra subtractions were applied. First, only instrumental

background was subtracted from the catalyst spectra acquired. This method was used in

Figure 4.3. Second, dehydrated catalyst spectra at 100 °C was subtracted from the spectra

of the adsorbed reactants at different temperature. This method was used in Figure 4.5.

Third, in temperature programmed desorption DRIFTS experiments, dehydrated catalyst

spectra at the exact same temperature were subtracted from the spectra of the adsorbed

reactant. This method was used in Figures 4.4-7. Fourth, in temperature programmed

desorption DRIFTS experiments, in the presence of the vapor reactant, a spectrum

containing catalyst and adsorbed surface species at 373 K was subtracted from the spectra

Page 84: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

64

containing contributions of the catalyst, adsorbed surface species and the vapor-phase at

specific temperature. This method was used in Figures 4.9-11.

Temperature-programmed reaction spectroscopy (TPRS) was carried out using an

Altamira Instruments system (AMI-200) connected to Dymaxion Dycor mass spectrometer

(DME200MS). Approximately 30 mg of catalyst was loaded into a glass U-tube fixed-bed

reactor and held in place by quartz wool. Prior to measurement, the catalyst was first pre-

treated under 10% O2/Ar (Airgas, certified, 9.99% O2/Ar balance) at 500°C for 1 hour.

After pretreatment, the catalyst temperature was brought down to 100°C. At this

temperature, ethanol was preadsorbed for 15 minutes, followed by degassing using argon

for 45 minutes. The vapor delivery system followed that of in-situ DRIFTS study. Finally,

the fixed-bed reactor was heated at ~10°C/min to 450°C in the flowing reactant gases and

the evolution of the products was monitored with the online mass spectrometer.

Acetaldehyde was delivered using a mixture of 5% acetaldehyde in nitrogen (Praxair).

Another experiment involved preadsorbing acetaldehyde on the surface of catalyst,

followed by degassing with argon for 45 minutes, and temperature increase at 10°C/min to

450°C under constant ethanol/argon flow. The ethanol saturator is chilled at 2°C using ice

bath at all times.

3.5. Operando XANES and EXAFS spectroscopy during ethanol reaction to 1,3-

BD over Cu- and Zn-promoted MgO/SiO2 catalysts

Operando X-ray absorption spectroscopy (XAS) was performed at the beamline BL2-2 at

Stanford Synchrotron Radiation Lightsource (SSRL), SLAC National Accelerator

Laboratory. The Cu and Zn K-edge data were collected in transmission mode. For the

measurements, the sample powder was loaded into a quartz tube with 0.9 mm inner

Page 85: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

65

diameter and 1.0 mm outer diameter which was then mounted into the Clausen plug-flow

reaction cell.54 Ethanol vapor was delivered into the system using a temperature-controlled

saturator to manipulate the space velocity. He was bubbled through the saturator and fed

into the reactor. Prior to the spectroscopy study under reaction condition, the catalyst was

pretreated at 450 °C for 1 hour under constant He flow. The operando measurements were

done at 100, 200, 300 and 400 °C under constant ethanol flow. After reactor temperature

reached 400 °C, the system was allowed to equilibrate for 2 hours and XAS spectra were

repeatedly taken. The operando condition was ensured by allowing the vapor-phase into a

dedicated RGA Mass Spectrometer (RGA, Stanford research system). Standard reference

compounds, CuO (Alfa Aesar), ZnO (Alfa Aesar), Cu2O (Alfa Aesar) and synthesized

reference materials, i.e. CuMg, ZnMg, CuSi, and ZnSi, were pressed into the pellets and

measured under ambient conditions.

Page 86: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

66

References

(1) Kresse, G.; Hafner, J. Phys. Rev. B 1994, 49 (20), 14251–14269.

(2) Kresse, G.; Furthmüller, J. Comput. Mater. Sci. 1996, 6 (1), 15–50.

(3) Kresse, G.; Furthmüller, J. Phys. Rev. B 1996, 54 (16), 11169–11186.

(4) Kresse, G.; Hafner, J. Phys. Rev. B 1993, 48 (17), 13115–13118.

(5) Blöchl, P. E. Phys. Rev. B 1994, 50 (24), 17953–17979.

(6) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77 (18), 3865–3868.

(7) Henkelman, G.; Jónsson, H. J. Chem. Phys. 1999, 111 (15), 7010–7022.

(8) Heyden, A.; Bell, A. T.; Keil, F. J. J. Chem. Phys. 2005, 123 (22), 224101/1-

224101/14.

(9) Fukui, K. J. Phys. Chem. 1970, 74 (23), 4161–4163.

(10) Fukui, K. Acc. Chem. Res. 1981, 14 (12), 363–368.

(11) Hratchian, H. P.; Schlegel, H. B. J. Phys. Chem. A 2001, 106 (1), 165–169.

(12) Jensen, F. Introduction to Computational Chemistry, 2nd ed.; John Wiley & Sons

Ltd., 2006.

(13) Kvisle, S.; Aguero, A.; Sneeden, R. P. A. Appl. Catal. 1988, 43 (1), 117–131.

(14) Angelici, C.; Weckhuysen, B. M.; Bruijnincx, P. C. a. ChemSusChem 2013, 6 (9),

1595–1614.

(15) Larina, O.; Kyriienko, P.; Soloviev, S. Catal. Letters 2015, 1–7.

(16) Zhang, M.; Gao, M.; Chen, J.; Yu, Y. RSC Adv. 2015, 5 (33), 25959–25966.

(17) Janssens, W.; Makshina, E. V.; Vanelderen, P.; De Clippel, F.; Houthoofd, K.;

Kerkhofs, S.; Martens, J. A.; Jacobs, P. A.; Sels, B. F. ChemSusChem 2014, 8 (6),

994–1008.

(18) Angelici, C.; Meirer, F.; van der Eerden, A. M. J.; Schaink, H. L.; Goryachev, A.;

Hofmann, J. P.; Hensen, E. J. M.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ACS

Catal. 2015, 5 (10), 6005–6015.

(19) Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A.

ChemSusChem 2014, 7 (9), 2505–2515.

(20) Che, M.; Tench, A. J. Adv. Catal. 1982, 31, 77–133.

(21) Sushko, P. V.; Shluger, A. L.; Catlow, C. R. A. Surf. Sci. 2000, 450 (3), 153–170.

(22) Kantorovich, L. N.; Holender, J. M.; Gillan, M. J. Surf. Sci. 1995, 343 (3), 221–

239.

(23) Chizallet, C.; Petitjean, H.; Costentin, G.; Lauron-Pernot, H.; Maquet, J.;

Bonhomme, C.; Che, M. J. Catal. 2009, 268 (1), 175–179.

(24) Chizallet, C.; Bailly, M. L.; Costentin, G.; Lauron-Pernot, H.; Krafft, J. M.; Bazin,

P.; Saussey, J.; Che, M. Catal. Today 2006, 116 (2), 196–205.

(25) Di Valentin, C.; Del Vitto, A.; Pacchioni, G.; Abbet, S.; Wörz, A. S.; Judai, K.;

Heiz, U. J. Phys. Chem. B 2002, 106 (46), 11961–11969.

(26) Branda, M. M.; Rodríguez, A. H.; Belelli, P. G.; Castellani, N. J. Surf. Sci. 2009,

603 (8), 1093–1098.

(27) Bailly, M.; Chizallet, C.; Costentin, G.; Krafft, J.; Lauronpernot, H.; Che, M. J.

Catal. 2005, 235 (2), 413–422.

(28) Baltrusaitis, J.; Hatch, C.; Orlando, R. J. Phys. Chem. A 2012, 116 (30), 7950–

7958.

(29) Haas, P.; Tran, F.; Blaha, P. Phys. Rev. B 2009, 79 (8), 85104.

Page 87: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

67

(30) Wu, C.-M.; Baltrusaitis, J.; Gillan, E. G.; Grassian, V. H. J. Phys. Chem. C 2011,

115 (20), 10164–10172.

(31) Badlani, M.; Wachs, I. E. Catal. Letters 2001, 75 (3–4), 137–149.

Page 88: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

68

Chapter 3

Computational Study of Ethanol to 1,3-BD

Reaction Mechanisms

Abstract ........................................................................................................................... 68

1. Introduction ........................................................................................................ 69

2. Computational Results ....................................................................................... 73

2.1. Reaction Pathways .................................................................................. 73

2.1.1. Ethanol dehydrogenation and dehydration ............................................. 73

2.1.2. Aldol condensation ................................................................................. 80

2.1.3. Prins condensation .................................................................................. 85

2.1.4. 1-Ethoxyethanol formation ..................................................................... 88

2.2. Details of the free-energy profiles .......................................................... 89

2.2.1. Elimination/redox reaction of ethanol .................................................... 90

2.2.2. C-C bond formation ................................................................................ 90

2.2.3. Proton transfer ......................................................................................... 91

3. Discussion ............................................................................................................ 92

4. Conclusion .......................................................................................................... 98

References ...................................................................................................................... 100

Abstract

In this work, we performed periodic Density Functional Theory calculations and

explored reactive pathways of ethanol catalysis to catalytically form 1,3-butadiene on

undoped MgO surface. We have identified critical reactive intermediates, as well as

thermodynamic and kinetic barriers involved in the overall reactive landscape. The overall

free energy surface was explored for the highly debated reaction mechanisms, including

Toussaint’s aldol condensation mechanism, Fripiat’s Prins mechanism and mechanism

based on Ostromislensky’s hemiacetal rearrangement. Thermodynamics and kinetics data

calculated showed four rate limiting steps in the overall process. In particular, ethanol

Page 89: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

69

dehydration to form ethylene possessed lower energy barrier than dehydrogenation to yield

acetaldehyde suggesting competing reactive pathways. C-C bond coupling to form

acetaldol (3-hydroxybutanal) is preceded with 16 kcal/mol forward reaction barrier. Direct

reaction of ethylene and acetaldehyde proceeds with a free energy barrier of 29 kcal/mol

suggesting that Prins condensation is an alternative route. Finally, thermodynamic stability

of 1-ethoxyethanol prevents further reaction via hemiacetal rearrangement. The results

here provide a first glimpse into the overall 1,3-butadiene formation mechanism on

undoped MgO reactive sites in light of the vast literature discussing variety of the proposed

mechanistic pathways mostly based on conventional homogenous organic chemistry

reactions.

1. Introduction

Since the rapid expansion of coal industry in the 18th century World has relied on

non-renewable sources for organic chemicals 1. Currently, petroleum and natural gas are

the main feedstocks of relatively inexpensive carbon source 2. Biomass can serve as a

sustainable and renewable carbon source to generate chemicals and bio derived ethanol

catalytic upgrading has been proposed as a viable route for biomass valorization 3,4. In

particular, ethylene, propylene, ethyl acetate, n-butanol and isobutene are some of the high

value chemicals that can be derived from ethanol 5–14. Furthermore, 1,3-butadiene, the most

important monomer for synthetic rubber, has been produced via catalytic processing of

ethanol during World War II in USSR and USA, using Lebedev and Ostromislensky

processes, respectively 15. The former utilized catalytic conversion of ethanol to 1,3-

Page 90: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

70

butadiene via one-step on MgO/SiO2 catalysts 16, while the latter utilized a two-step process

with the first step ethanol dehydrogenation to acetaldehyde over Cu/SiO2 catalysts 3,17

followed by acetaldehyde and ethanol coupling to 1,3-butadiene over a tantalum-based

catalyst. Catalysts for the one-step process were reported to have achieved ~50-60% yield

18,19, while the two-step process could attain over 60% yield, with purity of about 98-99%

at 300-350°C. Recent abundance of shale gas resulted in a different catalytic cracker

product distribution dominated by ethylene 3. This caused a worldwide shortage of C4

hydrocarbons, such as 1,3-butadiene. Since ethanol can be produced using variety of

biomass sources including fermentation and gasification, it recently emerged as the green

route to catalytically form 1,3-butadiene 3,4.

The biggest obstacle in ethanol catalysis to form 1,3-butadiene is relatively low

selectivity and the resulting yields of the desired product. Angelici et al. reported 74%

conversion with 49% selectivity on CuO/MgO-SiO2, while Makshina et al. described a

similar catalyst that attained 97.5% conversion with 58.2% selectivity 18,20. Most recently,

a CuO/HfO/ZnO catalyst was reported to have achieved 99% conversion with 71.1%

selectivity, e.g. ~70% yield 19. In general, doped-MgO supported on silica 18,20–29

(Lebedev’s catalyst) or mixed (supported) oxides 16,19,30–36 were used. Various transition

metal dopants (Zn, Cu, and Ag) were used to improve MgO/SiO2 catalyst performance, as

well as different synthesis methods and composition of MgO/SiO2 were investigated, such

as Kvisle’s wet kneading method 26 and the utilization of clay and sepiolite as the support

21,28. However, multitude of byproducts, including ethylene, C4 oxygenates and olefins,

diethyl ether, acetaldehyde and even acetone are still detected implying high separation.

This lack of the kinetic control over the ethanol-to-1,3-butadiene catalytic process and poor

Page 91: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

71

understanding of the fundamental mechanistic steps involved have hindered the

development of catalysts with reasonable performance. The generally accepted one-step

catalytic mechanism involves dehydrogenation of ethanol to acetaldehyde which then

undergoes C-C coupling via aldol condensation mechanism to yield crotonaldehyde.

Crotonaldehyde is further hydrogenated via MPV (Meerwein-Ponndorf–Verley) reduction

with ethanol and the resulting crotyl alcohol is dehydrated give butadiene37,38 as shown in

Figure 3.1a. In addition, Fripiat and Ostromislensky proposed two other possible reaction

pathways 39,40. Fripiat suggested Prins-like mechanism involving both dehydration and

dehydrogenation reactions producing ethylene and acetaldehyde, as shown in Figure 3.1b.

The C=O group is hydroxylated in the presence of Brønsted acid and reacts with ethylene

opening the double bond. The resulting 3-buten-2-ol is then dehydrated to yield 1,3-

butadiene 40. Ostromislensky’s version of the reaction mechanism shown in Figure 3.1c

involves the hemiacetal rearrangement between ethanol and acetaldehyde to yield 1-

ethoxyethanol that later converts to butane-1,3-diol 3. Two computational studies by

Chieregato et al. 41 and Zhang et al. 42 attempted to unravel the overall reaction mechanism.

Zhang et al. performed calculations using density functional generalized gradient

approximation (GGA) and focused on the very first step of the mechanism, e.g.

dehydrogenation of the alcohol. Stepped MgO surface was predicted to have lower energy

barrier than flat surface for this reaction 42. Chieregato et al., on the other hand, proposed

an entirely different mechanism using cluster type calculations and Gaussian basis set.

They ruled out crotonaldehyde and crotyl alcohol as possible intermediates and concluded

that acetaldehyde would react with a carbanion resulting from ethanol C-H cleavage. In

this work we performed ethanol catalytic coupling to 1,3-butadiene using a kink Mg atom

Page 92: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

72

at a step-edge MgO (100) as a model catalyst surface in accordance with the recent works

that suggest MgO as a bifunctional catalyst 43–53. The energetics and structure of key

reactive intermediates, e.g. acetaldehyde, crotonaldehyde, and crotyl alcohol, for

Lebedev’s reaction 16 based on the proposed mechanism by Toussaint et al. 38,54, as well as

the proposed Fripiat’s Prins and Ostromislensky mechanistic pathways were explored to

determine the kinetic limitations of ethanol catalytic coupling to 1,3-butadiene on MgO.

Figure 3.1. Reaction mechanisms proposed for ethanol to 1,3-butadiene; (a) Toussaint’s

generally accepted mechanism, (b) Fripiat’s Prins mechanism, (c) Ostromislensky’s

hemiacetal rearrangement.

Page 93: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

73

2. Computational results

2.1. Reaction Pathways

2.1.1. Ethanol Dehydrogenation and Dehydration

We begin with the first step of ethanol catalytic transformation into acetaldehyde

or into ethylene for which the computed free-energy profiles are shown in Figure 3.3.

States 1A-1C and 5A-5C in Figure 3.2 demonstrate the reaction pathways for ethanol

dehydrogenation and dehydration to form acetaldehyde and ethylene, respectively. The

corresponding TSs are labelled as 1B(TS) and 5B(TS). In both cases the reaction starts

from the structures formed upon spontaneous dissociative adsorption of ethanol whereby

proton is abstracted either by the edge or terrace oxygen atoms. The calculated relative free

energies for the configurations 1A and 5A are -13.5 and -10.5 kcal/mol, respectively, at

450 oC. In 1B TS, the surface bound proton becomes coordinated to the proton leaving -

carbon atom. The H….H distance is 0.84 Å, which is 0.07 Å longer than the equilibrium

H-H bond distance in hydrogen molecule in 1C. Compared to 1A, the distance between the

hydrogen atom and lattice O4C has significantly increased from 0.98 Å to 1.61 Å.

Furthermore, acetaldehyde is coordinated to Mg3c via oxygen atom while also accepting

some electron density from in the terrace O5c resulting in a distorted chemisorbed structure.

During the dehydration step hydrogen atom is adsorbed on terrace oxygen atom as shown

in Fig. 3.2 (5A). During the reaction over transition state 5B one of the β-hydrogen atoms

becomes oriented towards O4C with C-H and O4C-H of 1.45 and 1.22 Å, respectively. The

final state 5C results in two hydroxyl groups formed on the MgO surface with ethylene

molecule loosely coordinated to the surface. Relative electronic energies at T=0 K and the

corresponding free energy values at 450 oC for the reaction profile shown in Figure 3.3 are

Page 94: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

74

provided in Table 3.1. In particular, at 450°C the forward free-energy barrier for the

reaction 1A 1C is 39.6 kcal/mol while the barrier for the reverse barrier is only 20.5

kcal/mol suggesting that the acetaldehyde formation is endergonic. Ethylene formation via

5A 5C has barrier 33.5 kcal/mol but the reverse barrier is higher with 38.3 kcal/mol.

While both acetaldehyde and ethylene are typically observed as reaction byproducts of

ethanol catalytic coupling 18,20,22–25, the stability of ethylene vs acetaldehyde is intriguing

but not surprising. Ethanol is known to undergo intramolecular dehydration in the presence

of acidic and basic surface sites 55, while acetaldehyde formation in general needs redox

metals 56. Hence for undoped MgO catalyst ethylene generation is expected and preferred

over acetaldehyde.

The reaction pathways of ethanol dehydration and dehydrogenation products

further proceed via two main reaction mechanisms discussed in detail in Sec.3.1.2 and

3.1.3: aldol condensation and Prins condensation reaction. The aldol condensation pathway

entails acetaldehyde transformation into its enolate form followed by the reaction with the

molecular acetaldehyde to form a C-C bond. The resulting C4 intermediate then undergoes

several steps of intermolecular proton transfer with the surface and ethanol to yield 1,3-

butadiene. Prins condensation entails C-C bond formation via reaction of acetaldehyde

and ethylene followed by proton transfer steps. In these mechanisms, the proton diffusion

through the surface leads to water release from the surface.

Page 95: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

75

Table 3.1 Electronic and free energy values of the stationary points calculated at 0 K and

723 K, respectively.

State

Electronic

energy

(kcal/mol)

Referenced

free energy

(kcal/mol)

State

Electronic

energy

(kcal/mol)

Referenced

free energy

(kcal/mol)

1A -40.6 -13.5 4F -45.3 -8.1

1B 3.9 26.2 4G -43.2 -6.1

1C -16.6 5.6 4H -52.0 -46.2

2A -14.3 -8.5 4I -37.8 -37.7

2B 4.1 7.4 4J -5.5 -5.3

2C -22.2 -20.3 4K -18.7 -24.0

2D -17.7 -17.3 5A -39.9 -10.5

2E -26.2 -28.8 5B -2.6 23.0

2F -22.5 -23.7 5C -32.0 -15.3

2G -15.9 -7.6 6A 11.6 -0.2

2H -19.4 -10.8 6B 25.6 28.6

2I -12.5 -1.6 6C -34.8 -22.9

2J -11.4 -1.6 6D 3.7 15.8

2K -6.2 4.2 6E -26.8 -18.9

2L -15.4 -8.2 6F 9.5 15.3

2M 9.3 15.7 6G -6.0 -0.7

2N 2.2 3.2 6H -4.4 -2.9

2O 6.8 -18.2 6E i -21.8 -16.0

3A -24.1 -18.8 6E ii -30.5 -27.2

3B -14.6 -6.8 6E iii -32.5 -27.3

3C -19.8 -15.2 6E iii 1 2.2 5.6

3D -27.3 -27.8 6E iii 2 -13.8 -9.6

3E -1.3 -0.2 6E iii 3 -11.5 -12.7

3F -15.4 -15.4 6E iv 43.9 44.3

3G -15.6 -20.4 6E v -21.4 -25.4

4A -53.1 -14.6 7A -51.2 -26.4

4B -52.5 -4.8 7B -30.6 2.5

4C -64.4 -22.1 7C -33.4 3.9

4D -68.3 -32.0 7D -30.5 6.4

4E -42.5 -5.8 7E -44.8 -8.7

Page 96: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

76

1A 1B (TS) 5A 5B (TS)

1C 5C

2A 2B (TS) 2C 2D (TS)

2E 2F 2G (TS) 2H

2I 2J 2K (TS) 2L

Page 97: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

77

2M (TS) 2N 2O

3A 3B (TS) 3C 3D

3E (TS) 3F 3G 4A

4B (TS) 4C 4D 4E (TS)

4F 4G (TS) 4H 4I

Page 98: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

78

4J (TS) 4K

6A 6B (TS) 6C 6D (TS)

6E 6F (TS) 6G 6H

6E i (TS) 6E ii 6E iii 6E iv (TS)

6E v 6E iii 1 (TS) 6E iii 2 6E iii 3

Page 99: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

79

7A 7B (TS) 7C 7D (TS)

7E

Figure 3.3 Free-energy profiles for (a) ethanol dehydrogenation to form acetaldehyde

and (b) ethanol dehydration to ethylene.

(a) (b)

Figure 3.2 All stable intermediates and transition states calculated following the reaction

pathways. (1A-1C): ethanol dehydrogenation to acetaldehyde; (2A-2O): acetaldehyde

aldol condensation to 3-hydroxybutanal (acetaldol) followed by proton transfer to

crotonaldehyde; (3A-3G): MPV (Meerwein–Ponndorf–Verley) reduction of

crotonaldehyde to 1,3-butadiene; (4A-4K): acetaldol MPV reduction to butadiene; (5A-

5C): ethanol dehydration to ethylene; (6A-6E iii 3): Prins condensation of acetaldehyde

and ethylene; (7A-7E): ethanol and acetaldehyde nucleophilic addition reaction

(Ostromislensky’s hemiacetal rearrangement).

Page 100: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

80

2.1.2. Aldol condensation

Classical aldol condensation mechanism requires one of the acetaldehyde

molecules to be in its enolate state.57 In this work the enolate state 2C (see Fig. 3.2) was

obtained via proton transfer of β-hydrogen to terrace atom O5C yielding a hydroxyl group

via low energy barrier 2B TS from the initial stable strongly adsorbed acetaldehyde

molecule in 2A with free-energy of -8.5 kcal/mol relative to the reference state. The

forward barrier for the reaction step 2A 2C is 16 kcal/mol as shown in Figure 3.4. This

mechanism is facilitated by the C=O bond elongated from 1.21 Å58 to 1.43 Å due to strong

interaction with the surface oxygen atoms. In TS configuration, one of hydrogen atoms

from the methyl group establishes a hydrogen bond with surface oxygen atom causing an

elongation of the corresponding C-H bond from 1.10 to 1.29 Å. State 2C represents a stable

configuration with sp2 hybridized carbon enolate atoms and surface hydroxyl group.

For the aldol condensation to take place, the hydrogen atom bound to the surface

needs to be in a close proximity to the enolate molecule requiring it to diffuse to the edge

O4C atom. This transition proceeds via transition state structure 2D with the forward barrier

of only 3 kcal/mol. The next step is the physisorption of a second acetaldehyde molecule

on the surface in 2F, preceding the C-C bond formation via aldol condensation (2F-2H).

The TS for this step (the structure 2G) shows the coordination between enolate and

acetaldehyde with the reactive β-carbon of enolate and the -carbon of acetaldehyde

adsorbed on surface site Mg3C establish a C-C bond of length of 2.07 Å. The forward IRC

analysis for this reaction shows the formation of C-C bond between the two reactive carbon

atoms. The length of the latter bond in the stable structure 2H is 1.64 Å. Formation of the

Page 101: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

81

acetaldol (3-hydroxybutanal) in 2H is preceded with forward reaction barrier of 16.1

kcal/mol.

Once 3-hydroxybutanal is formed, the aldol needs to lose a hydrogen atom to the

surface to undergo dehydration to yield crotonaldehyde. For this step to take place, a

reactive O4C is required. Assuming transient proton diffusion between O4C and O5C atoms,

proton abstraction takes place followed by the transformation of the trans-isomer, as

depicted in 2I, into 2J via aergonic steps. The activation energy from cis- to trans-

crotonaldehyde reported is ~13 kcal/mol 59. The molecule subsequently loses hydrogen

Figure 3.4 Free-energy profiles for aldol condensation

pathway.

Page 102: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

82

atoms to the MgO surface via step 2J2L (ΔGA,forward = 5.9 kcal/mol) and desorbs from

the surface after breaking the C-O bond via sequence 2L2M(TS)2O (ΔGA,forward = 24

kcal/mol) yielding the structure 2O with a newly formed surface site O3c which under

reactive conditions can recombine with protons to form H2O reforming the original Mg3c

site.

Formation and desorption of crotonaldehyde in 2N-2O agrees very well with the

occasional gas phase byproduct observations 22,37. Formation of the O3C surface site in 2O

is followed by the water molecule formation, which is another product of ethanol coupling

reactions. The next step in the overall mechanism is the MPV reduction of crotonaldehyde

by ethanol (3A3C) to form adsorbed acetaldehyde in 3C followed by its desorption and

proton transfer to the surface (3C3G) to form 1,3-butadiene. The corresponding free-

energy profile (2H 2O 3A 3G) is shown in Figure 3.5. The highest barrier that

we determined within this sequence was that of proton transfer to the surface in 3E TS with

27.6 kcal/mol. Reduction of the unsaturated aldehyde by hydrogen was assumed not to take

place as confirmed by the measured hydrogen content in the reaction products 30 and due

to the gas-phase thermodynamic calculation which favors the reduction by ethanol 3.

Dissociation of hydrogen on the defected MgO surface itself is a non-spontaneous process

with a relatively low activation barrier of 2.8 kcal/mol 60 whereas ethanol dissociation on

Mg3c is spontaneous as shown in Figure 3.2. Additionally, the heterolytic dissociation can

only be stabilized on a high density of 3-coordinated sites which suggests small amount of

surface hydrogen 61.

Page 103: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

83

Page 104: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

84

Alternatively, the 1,3-butadiene formation mechanism proposed earlier by

Ostromislensky (also vide infra) involves the hemiacetal rearrangement 39. It was argued

that ethanol can react with acetaldehyde to form 1-ethoxyethanol which will then undergo

rearrangement to butane-1,3-diol and further dehydrate to 1,3-butadiene. However, this

mechanism has been rebuffed by Quattlebaum et al. 37. The formed C-O-C bond, if it is to

be rearranged to make C-C bond, would lead to its dissociation. The identified butane-1,3-

diol, however, could be formed when acetaldol is reduced by ethanol, as shown in Figure

3.5 4A-4C. State 4C is effectively dissociated (adsorbed) butane-1,3-diol which is formed

via series of exergonic steps 2J-4A-4B TS-4C with a very low forward free-energy barrier

of 9.8 kcal/mol. The resulting adsorbed butane-1,3-diol can further undergo several steps

of proton transfer to yield 1,3-butadiene (4C-4K). In this situation, there are three

competing processes with transition states 4E, 4G, and 4J. The reaction channel with TS

4G breaks a C-O bond of the adsorbed butane-1,3-diol. The reaction 4F4H is extremely

Figure 3.5 Free-energy profiles for the MPV reduction of the resulting molecule from

aldol condensation. Red pathway indicates subsequent proton transfer of acetaldol

followed by MPV reduction of the crotonaldehyde; Blue pathway shows the direct MPV

reduction of the resulting acetaldol.

Page 105: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

85

exergonic with very low free-energy of activation (~2 kcal/mol). It is preceded by the 4E

TS and the free-energy of activation of 26.2 kcal/mol, which is a typical value to that of

sp3 proton transfer to the reactive surface O4c atoms. The last step of this condensation

mechanism is the simultaneous C-O bond breaking and proton transfer to the surface

(4I4K). Fig. 3.2 (4J) depicts a transition state where the 1,3-butadiene is desorbing from

the surface. Interestingly, the two different MPV reduction steps yield two different

conformations of 1,3-butadiene. MPV reduction of crotonaldehyde gives s-trans

conformation, while that of acetaldol results in s-cis conformation (structures 3G and 4K,

respectively, see Fig. 3.2). The stable conformation is, however, trans 1,3-butadiene, which

makes an additional step for acetaldol reduction necessary. This last step will be cis/trans

isomerization to trans 1,3-butadiene with rather low free-energy of activation of only ~4

kcal/mol 62.

2.1.3. Prins condensation

Our data shown in Figure 3.2 suggest that on undoped MgO the ethylene formation

from ethanol will compete with that of acetaldehyde. Prins condensation is among the early

proposed mechanisms for ethanol reaction to 1,3-butadiene.40 The explicit ethanol reaction

mechanism on MgO via Prins mechanism is studied in this work and the corresponding

results are presented in this section. The corresponding structures are shown in Figure 3.2

(6A-6E) whereas the free-energy reaction profiles are displayed in Figure 3.6. The Prins

condensation pathway is the formation of C-C bond by opening the double C=C and C=O

bonds of both ethylene and acetaldehyde, respectively (6A-6C). 6B TS represents the

double bond opening of C=O with the oxygen coordinated between two surface Mg atoms

(Mg4C and Mg3C). This charge transfer to the MgO surface makes the -carbon susceptible

Page 106: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

86

to attack by the sp2 carbon molecule. The intermolecular C-C distance is now 1.90 Å, while

that of the aldehyde C-O bond is elongated by 0.1 Å. The double bond opening results in

a C4 structure bound to the surface, as shown in Fig. 3.2 (6C). The corresponding forward

free-energy barrier is 28.8 kcal/mol. This C-C coupling step is then followed by the proton

transfer to the surface atom O5c and the simultaneous C-Olattice bond breaking (see Fig. 3.2

(6C-6E)) followed by another proton transfer from the terminal sp3 carbon to the O4c

surface atom (see Fig. 3.2 (6E-6G)). The free-energy barriers for these steps are of 38.7

and 34.1 kcal/mol, which are typical values for proton transfer reactions considered in this

work. The structure 6H represents the desorbed structure of 1,3-butadiene.

An alternative pathway for transformation of the structure 6E is that via transition

state 6E i leading to the product 6E ii (see Fig. 3.2). Here instead of the subsequent proton

transfer from the terminal carbon (as in the reaction 6E6G) the surface proton diffuses

from planar surface atom O5c to a nearby edge atom O4c via low forward barrier of 2.9

kcal/mol. The terminal proton transfer then takes place, as depicted in Fig. 3.2 (6E iii, 6E

iii 1, 6iii 2) and the computed free-energy of activation (32.9 kcal/mol) is comparable to

that for steps alternative pathway (6C6E and 6E6G). Interestingly, a cyclic TS can

also be established in another variant of this mechanism, see Fig. 3.2 (6E iv). This

mechanism yields a physisorbed molecule as a product, which is similar to methylethyl

ketone (MEK) shown in Fig. 3.2 (6E v). However, this step has rather large forward

activation energy of 71.6 kcal/mol.

Page 107: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

87

Figure 3.6 Free-energy profiles for the Prins condensation between ethylene and

acetaldehyde. Red pathway indicates a typical route of Prins condensation; Blue

pathway shows an additional proton diffusion step in between the reaction steps; Black

pathway shows the unlikely formation of MEK.

Page 108: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

88

2.1.4. 1-ethoxyethanol formation

Final major reaction mechanism considered in this study is the Ostromislensky’s

hemiacetal rearrangement which will be discussed in this section. The very first step in this

case is the reaction of ethanol and acetaldehyde to yield 1-ethoxyethanol, which was further

postulated to undergo a molecular rearrangement to form butane-1,3-diol. The very first

step was investigated and it was found to proceed via stationary structures 7A to 7E shown

in Figure 3.2. The free-energy profile for these steps is shown in Figure 3.7. The initial

structure 7A contains an ethoxy species formed during the chemisorption of ethanol, as

well as a molecule of acetaldehyde physisorbed to the MgO surface. The C-O bond

formation to yield 1-ethoxyethanol (via transition state 7B) has a forward barrier of 28.9

kcal/mol with several nearly isoexergonic molecular rearrangements followed by the

stabilized 7E structure with former aldehyde C-O that is still coordinated to the surface

atom Mg3c. It is apparent from Figure 3.7 that reverse barrier for the 1-ethoxyethanol

formation is almost zero. In this case 1-ethoxyethanol can behave as a thermodynamical

sink that would form in a transient fashion before reacting via other discussed pathways to

form 1,3-butadiene. Surprisingly, the free energy computed for the structure 7C, which is

a minimum on potential energy surface (PES), is slightly higher than that for the structure

7B – TS that is a first-order saddle point on PES. This unexpected result is clearly due to a

failure of the harmonic approximation used in this work to determine free-energies. This

level of theory implies that the positions of stationary points on PES and on the free-energy

surfaces are identical which is generally not true (see Ref.63 for discussion of the limitations

of harmonic transition state theory). Furthermore, this level of theory is unsuitable to

describe soft degrees of freedom such as hindered molecular rotations or long-wave lattice

Page 109: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

89

vibrations which contribute to harmonic free-energies more than the hard ones. In our case,

the free energy for the first-order saddle point structure 7B is 28.9 kcal/mol higher than

that for the minimum 7A and, importantly, the reaction coordinate for the whole sequence

7A→7E consists of hindered rotations of the CH3CH2O- and CH3CHO- groups with the

imaginary frequencies for the TS structures 7B and 7D that are smaller than 100 cm-1. We

note, however, that the free-energies for the sequence of structures between two stable

configurations 7A and 7E (i.e. the structures 7B, 7C and 7D) are all within 4 kcal/mol and

this number is relatively small compared to the free-energy difference with respect to the

stable structure 7A. As the configuration 7D is the one with the highest free-energy on the

sequence of steps 7A→7E, we consider the difference G(7D)-G(7A)=32. 8 kcal/mol as the

effective free-energy barrier for the whole process 7A→7E.

2.2. Details of the Free-energy profiles

Three particular steps will be discussed here, namely elimination/redox reactions of

ethanol, C-C bond formation, and proton transfer.

Figure 3.7 Free-energy profile for ethanol and acetaldehyde nucleophilic addition

reaction.

Page 110: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

90

2.2.1. Elimination/redox reaction of ethanol

Elimination reaction takes place when a substituent leaves the molecule, e.g. water leaving

ethanol, while redox reaction is defined as a reaction where a molecule loses or gains

hydrogen. 64 Dehydrogenation reaction of ethanol (redox) is the first and foremost reaction

step in all mechanisms proposed for the 1,3-butadiene formation. This oxidation step yields

hydrogen as a byproduct while also transforming ethanol into acetaldehyde, a more reactive

intermediate. The transformation 1A1C shows a rather high free-energy barrier, 39.6

kcal/mol, while the reaction itself is endergonic in nature, with ΔGRx=20.5 kcal/mol. On

the other hand, ethanol dehydration to ethylene has slightly lower activation barrier, 33.5

kcal/mol, and it is slightly exergonic with ΔGRx=-4.9 kcal/mol. Comparison of both

reactions shows that ethanol is more likely to lose water than hydrogen on undoped MgO

catalyst surface, which means that ethylene should be produced in higher amounts than

acetaldehyde. This is in agreement with the experimental reports where small amounts of

Ag, Cu or Zn are typically incorporated into the lattice structure to enhance acetaldehyde

formation 23,65–71.

2.2.2. C-C bond formation

Two C-C bond formation pathways are presented in this study. Namely, aldol condensation

and Prins condensation. The pathway 2F2H possesses a favorable activation energy,

lower than that of Prins, 16.1 v 28.8 kcal/mol. However, this reaction is thermodynamically

limited, as shown by the endergonic nature of the reaction, with ΔGRx=12.9 kcal/mol. This

suggests that aldol condensation step is kinetically favored on undoped MgO samples, with

arguably one of the most favorable steps in the whole reaction landscape, while the

exergonic nature of the Prins mechanism makes it thermodynamically favored. The

Page 111: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

91

activation energy of the latter is also similar to the energetic barrier to other steps,

comparable to that of ethylene formation, and even lower than the activation energy of

ethanol dehydrogenation. However, the overall picture is more complex since ethylene

formed is only physisorbed onto the surface and adsorption of both reactants, i.e. ethylene

and acetaldehyde, is only -0.2 kcal/mol lower in free-energy than the reference state

suggesting that both molecules can desorb as byproducts. In accord with implications of

these results, both ethylene and acetaldehyde have been seen as byproducts of ethanol

catalytic coupling to form 1,3-butadiene.18,20,22–25

2.2.3. Proton transfer

Proton transfer steps can be further subdivided into three categories. The proton transfer

steps of the first category are those that take place between the organic molecule and the

surface, e.g. 2J2L, 3D3F, 4D4F, 4I4K, 6C6G. The second type of proton

transfer reactions is the MPV reduction taking place between two organic molecules, e.g.

3A3C and 4A4C. The last type is the proton diffusion from one site of the MgO

surface to another, e.g. 6E6E i6E ii, 2C2E, and also the understated proton diffusion

steps between 2H and 2I.

The first type of proton transfer typically exhibits moderate activation energy. Most of the

cases have ΔGA,forward = ~30 kcal/mol with only one case, i.e. 2J2L, possessing very low

activation energy of ~6 kcal/mol, possibly due to the very saturated nature of the organic

C4 compound. In the case of the MPV reduction, hydrogen atom moves from an alcohol

-carbon to open up the C=O bond of a molecule. This reduction reaction does not have a

typical activation energy but rather it depends on the nature of the C=O containing

molecule itself. The values computed for crotonaldehyde and acetaldol are ~12 kcal/mol

Page 112: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

92

and ~10 kcal/mol, respectively. Finally, the last type of the proton transfer reaction is

proton diffusion from one MgO surface site to another. This reaction has typically very

low activation energy of ~3 kcal/mol which suggests that the water formation and

desorption are easily facilitated by the MgO catalyst.

3. Discussion

Adsorption of ethanol on both perfect and defect sites of MgO surface had been

studied previously using cluster calculation 47. It was shown that ethanol dissociated on

defect sites but not on the perfect surface. Moreover, the adsorption energy decreased with

the coordination number of the adsorption site 47. This finding also aligns well with our

calculations which show two modes of ethanol dissociation on the defect sites, i.e. 1A and

5A. In the state 1A, adsorbed molecules are coordinated on Mg3C (corner) and O4C (edge),

while in the configuration 5A the ethanol molecule is chemisorbed on Mg3C and O5C

(terrace). The energy of the former is lower than the latter (both electronic energy and

Gibbs’ free energy), indicating the difference in stability of both states. The lower

coordination O4C is very reactive and hence the spontaneous chemisorption of ethanol 45,52.

Figure 3.2 structure 5A, however, depicts the adsorption of ethanol on the same Mg atoms

(Mg3C and Mg5C), but proton adsorbing on O5C. The highly-coordinated O5C does not

possess similar deprotonation ability to its lower-coordinated counterparts, as it is already

stabilized by coordination with the neighboring atoms 45,50. This situation causes the ethoxy

oxygen to interact less strongly with the surface in Figure 5A, resulting in a relatively more

unstable state compared to Figure 3.2 structure 1A. Similarly, two new hydroxyl groups

Page 113: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

93

are formed during the dehydration process and only weakly bound acetaldehyde and

molecular hydrogen during the dehydrogenation process in Figure 3.2 structures 5C and

1C, respectively, with the former being more stable.

A study by Zhang et al. showed a peculiar finding where ethanol dissociated on

both perfect and defected sites with energy barriers of 1.63, 1.42, and 1.30 eV, for terrace,

kink, and edge, respectively 42. Surprisingly, the molecule needed to surpass higher barrier

for kink which consisted of two low coordinated ions (Mg3C-O3C), than for stepped Mg4C-

O4C. This is in contrast to findings reported in this work and those of Branda where strong

dissociation on the defect sites takes place without any barrier 47. Finally, Chieregato et al.

41 showed that ethanol dehydrogenation over corner site of the MgO surface had an

energetic barrier of 44.7 kcal/mol on Mg3C site, as determined using cluster B3LYP/6-

31++G(d,p) DFT calculations. Furthermore, the reaction was also postulated to be slightly

exergonic with respect to the gas phase reference components, with ΔE of -1.4 kcal/mol.

Our periodic calculations, on the other hand, predict that ethanol to acetaldehyde has a

rather high energetic barrier, and it essentially represents rate-limiting step in the overall

mechanism. The free-energy barrier, based on our calculation was 39.6 kcal/mol at 450°C.

The ΔGRx for this reaction is also calculated to be +19.1 kcal/mol, which is highly

endergonic.

The free-energy values for the profiles presented in Figure 3.3-7 are listed in Table

3.2, along with the computed reaction rate constants. Based on the results presented in

Figure 3.3, ethylene is more likely to be produced than acetaldehyde due to the lower

activation energy and the exergonic nature of the reaction. As depicted in Figure 3.4 only

one TS structure has barrier higher than the desorption energy of the molecules in the

Page 114: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

94

reference state, namely that for enolate formation (2A-2C). In the subsequent step C-C

bond coupling takes place between the enolate and the physisorbed acetaldehyde (2F-2H).

This transition state (2G) is facilitated by enol, acetaldehyde as well as the resulting C-C

product bonded by low-coordinated Mg atoms.

The resulting acetaldol, after several steps of proton diffusion and isomerization to

cis conformer, can either lose proton to the surface or undergo MPV reduction, as shown

in Figure 3.5 for steps 2J2L and 2J4C, respectively. The two different pathways show

that MPV reduction of acetaldol can be more favorable with a sharp decrease in its energy

when ethanol is adsorbed, i.e. ethanol adsorption is much more favorable than a proton

transfer from acetaldol to the surface. One should note that the overall MPV reduction

pathway of acetaldol is below the reference state, which means that all the reaction steps

are more favorable than the desorption of any adsorbates. State 4D, which results from the

subsequent acetaldehyde desorption from state 4E, is essentially an adsorbed butane-1,3-

diol. This is the basis of Ostromislensky’s reaction mechanism supported by our

calculations, although the rearrangement step from 1-ethoxyethanol to this diol has been

previously rejected.39 As mentioned before, MPV reduction of acetaldol pathway would

require an additional step to convert the cis-1,3-butadiene to trans-1,3-butadiene (ΔGRx =

~4kcal/mol), which is the more stable molecule.

The adsorption of crotonaldehyde on the defected surface shows that the C=O bond

is now lengthened from 1.25 Å (gas-phase) to 1.27 Å. This bond lengthening, also noticed

by Boronat, et al., is attributed to the back-donation of the surface antibonding orbital

π*(CO) of crotonaldehyde 72. These authors calculated three main pathways for MPV

reduction of cycohexanone by 2-butanol over a tin-zeolite catalyst and reported the most

Page 115: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

95

favorable pathway (ΔGA,forward = ~15 kcal/mol) proceeding via formation of alkoxy species

on the surface, although their calculation was carried out on a single metal center model 72.

Furthermore, both direct MVP and H-transfer facilitated by metal hydride formation have

been reported with the former taking place over alkali-catalysts while the latter over

transition metal catalysts 73,74. A direct MPV reduction mechanism was also reported to

take place during the 5-HMF reduction by methanol on Mg3c site of MgO cluster model,

as reported by Pasini, et al. 73 with electronic energy of 27.5 kcal/mol.

Table 3.2 Computed forward and reverse reaction barriers and the corresponding reaction

rate constants.

Reaction ΔGA (kcal/mol) K (s-1)

Forward Reverse Forward Reverse

1A 1C 39.6 20.5 15.9 9.37 106

2A 2C 16.0 27.7 2.26 108 6.26 104

2C 2E 3.0 10.5 1.84 1012 9.93 109

2F 2H 16.1 3.2 2.04 108 1.60 1012

2J 2L 5.9 12.5 2.52 1011 2.59 109

2L 2N 24.0 12.5 8.66 105 2.44 109

3A 3C 12.1 8.5 3.35 109 4.17 1010

3D 3F 27.6 15.3 6.88 104 3.67 108

4A 4C 9.8 17.3 1.69 1010 8.74 107

4D 4F 26.2 2.4 1.77 105 2.86 1012

4F 4H 2.1 40.2 3.57 1012 11.0

4I 4K 32.5 18.8 2.26 103 3.18 107

5A 5C 33.5 38.3 1.15 103 38.9

6A 6C 28.8 51.5 3.07 104 4.11 10-3

6C 6E 38.7 34.7 29.6 4.93 102

6E 6G 34.1 2.3 7.45 102 3.14 1012

6E 6E ii 2.9 11.3 2.00 1012 6.00 109

6E iii 6E v 71.6 69.7 3.43 10-9 1.26 10-8

6E iii 6E iii2 32.9 15.2 1.76 103 3.92 108

7A 7E 32.8 15.1 1.86 103 4.24 108

Page 116: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

96

Free-energy profile for the Prins pathway is depicted in Figure 3.6. Both

acetaldehyde and ethylene (6A) have rather low adsorption energies, i.e. the two C2 species

can easily desorb from the surface before going to the anticipated TS (6B). Notably, all

transition states in this mechanism have positive relative energy with respect to the

reference state. The blue pathway indicates a Prins mechanism that includes proton

diffusion which results in a slightly lower free energy of the last transition state (6E iii 1)

compared to the original red pathway, of which last transition state has a higher free energy

(6F). The final state in both variants of Prins mechanism is trans-1,3-butadiene detached

from the surface. Another step considered within the discussion of Prins mechanism is the

formation of highly energetically unfavorable cyclic TS (6E iv). Not only does it have a

large activation barrier, but it also goes to another minimum (6E v) which has a slightly

higher relative energy than the initial state (6E iii).

The Prins mechanism was originally suggested by Gruver et al. 40. The authors used

aluminated sepiolites (both ammonium-exchanged and silver-exchanged) for the butadiene

production from ethanol. On the silver exchanged catalyst, the production of ethylene and

1,3-butadiene increased exponentially with increasing contact time, while acetaldehyde

production was linear 40. The adoption of the same mechanism for MgO can be attributed

to the fact that both catalysts possess almost exclusively Lewis acid sites 75 as Prins is a

mechanism that mostly takes place on Lewis acid sites 76. The reason that this mechanism

has not been considered viable was the postulated step of ethylene protonation, which was

supposed to result in a highly unstable carbocation 3. As shown in this work, there is another

type of intermediate/transition state for Prins mechanism which does not require

protonation of ethylene. This intermediate was also identified by Yamabe, et al. 77. In their

Page 117: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

97

theoretical work, propylene and formaldehyde were reacted via a novel C4 intermediate.

Finally, other evidences of reaction between olefin and aldehyde were shown in US Patent

no 2377025 A for 1,3-butadiene production, albeit through an acetylene intermediate, on

alumina-silver and Cr- and Mo-oxide catalyst78, isobutylene and pentenes with

formaldehyde on KU-2 cation exchanged resin to make dioxanes 79 and 1,3-butanediol

production via reaction of propylene and formaldehyde over ceria catalysts that contain

mostly Lewis acid sites 80.

Work presented here for the enol formation step also shows a much lower free-

energy barrier, compared to ethanol dehydrogenation, with a much more negative ΔGRx of

reaction, 16 and -11.8 kcal/mol, respectively. Chieregato et al. suggested a novel

mechanism with ethanol releasing a proton from its β-carbon and yielding a carbanion with

~33-36 kcal/mol forward barrier and negligible reverse barrier. This carbanion would then

react with either ethanol or acetaldehyde to yield 1-butanol or crotyl alcohol, respectively,

which subsequently dehydrate to produce 1,3-butadiene41. The carbanion has an interesting

configuration in which the ethanol is not deprotonated, rather the hydroxyl group is

interacting with a proton detached from β-carbon. From their overall mechanism, the rate-

limiting step was predicted to be the reaction of acetaldehyde and carbanion with the

electronic energy barrier of 11.4 kcal/mol with respect to the adsorbed reactants. The

reaction to form C4 hydrocarbons is exergonic. Our attempt on cleaving the proton from

β-carbon, however, lead to another pathway, namely to dehydration to form ethylene in

5A, 5B TS and 5C in Figure 3.2 (ΔGA,forward =33.5 kcal/mol, ΔGRx=-4.9 kcal/mol). The

unstable carbanion situation that would lead to C-O bond scission was not encountered in

the case of diol transformation to 1,3-butadiene (Figure 3.2D-K). As a result, Figure 3.3F

Page 118: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

98

shows a C4 molecule with two oxygen atoms bound to the surface. The terminal carbon,

however, is in distorted sp2 configuration and thus, represents a carbanion. Similarly, on

the investigated Prins mechanism, a stabilized C4 carbanion, which leads to the desorption

of 1,3-butadiene, is also observed in Figure 3.5G and 3.5E iii 2.

Interesting observation stemming from our work was that the Prins mechanism for

C-C bond formation was thermodynamically more favorable than aldol C-C coupling step,

and with the calculated barrier of 28.8 kcal/mol, i.e. ~10 kcal/mol lower than ethanol

dehydrogenation. The activation energy is, however, still larger than that of aldol

condensation (16.1 kcal/mol). Another fact complicating our conclusions further is that

adsorption of both C2 intermediates on the surface is almost unfavorable

thermodynamically (adsorption free energy (ΔGAds) = -0.2 kcal/mol), and the transition

state is located above the reference state. This step was similar to the

carbanion/acetaldehyde reaction, which is also a double bond opening of two sp2 carbon

atoms in acetaldehyde and ethylene. The suggested carbanion/acetaldehyde reaction

involved butane-1,3-diol as an intermediate which subsequently deprotonates, as opposed

to but-3-en-2-ol (state 5C) suggested by our calculation. 1,3-butanediol, however, still

appears in our mechanism as a product of MPV reduction on the resulting acetaldol from

aldol condensation.

4. Conclusion

A complex reactive mechanism of ethanol to form 1,3-butadiene was explored

using periodic quantum chemical methods. Overall free energy surface was explored for

Page 119: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

99

the highly debated reaction mechanisms, including Toussaint’s aldol condensation

mechanism, Fripiat’s Prins mechanism and mechanism based on Ostromislensky’s

hemiacetal rearrangement. Based on the thermodynamic and kinetic data determined

within this study we identified four rate limiting steps in the overall process. In particular,

ethanol dehydration to form ethylene possessed lower energy barrier than dehydrogenation

to yield acetaldehyde suggesting competing reactive pathways. Aldol condensation step to

form acetaldol is preceded with forward free-energy barrier of 16.1 kcal/mol but limited

thermodynamically with endergonic reaction free energy of 12.9 kcal/mol. This calculation

also offers another viable route in the form of Prins condensation, which has a free energy

barrier of 28.8 kcal/mol with exergonic reaction free energy of -22.7 kcal/mol. Finally,

thermodynamic stability of 1-ethoxyethanol prevents further reaction via hemiacetal

rearrangement. The results presented here provide a first glimpse into the 1,3-butadiene

formation mechanism on undoped MgO reactive sites in light of the vast literature

discussing variety of the proposed mechanistic pathways mostly based on conventional

homogenous organic chemistry reactions. While the surface model employed in this work

utilized most reactive MgO site, presence of H2O as a reaction product suggests that other

surface sites, based on reactive hydroxyls, can also affect the overall reactive pathways and

will be the focus of the future studies. However, based on the present calculations alone

several mechanisms appear possible. Reactivity experiments are needed to discriminate

between the different hypothesis, and we hope that our calculations will stimulate such

studies.

Page 120: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

100

References

(1) Matar, S.; Hatch, L. F. Chemistry of Petrochemical Processes; Elsevier, 2001.

(2) McCoy, M.; Reisch, M. S.; Tullo, A. H.; Short, P. L.; Tremblay, J. F.; Storck, W.

J. Chem. Eng. News 2007, 85, 29.

(3) Angelici, C.; Weckhuysen, B. M.; Bruijnincx, P. C. a. ChemSusChem 2013, 6 (9),

1595–1614.

(4) Sun, J.; Wang, Y. ACS Catal. 2014, 4 (4), 1078–1090.

(5) Han, Y.; Lu, C.; Xu, D.; Zhang, Y.; Hu, Y.; Huang, H. Appl. Catal. A Gen. 2011,

396 (1–2), 8–13.

(6) Zhang, X.; Wang, R.; Yang, X.; Zhang, F. Microporous Mesoporous Mater. 2008,

116 (1–3), 210–215.

(7) Meng, T.; Mao, D.; Guo, Q.; Lu, G. Catal. Commun. 2012, 21 (0), 52–57.

(8) Song, Z.; Takahashi, A.; Nakamura, I.; Fujitani, T. Appl. Catal. A Gen. 2010, 384

(1–2), 201–205.

(9) Sun, J.; Zhu, K.; Gao, F.; Wang, C.; Liu, J.; Peden, C. H. F.; Wang, Y. J. Am.

Chem. Soc. 2011, 133 (29), 11096–11099.

(10) Liu, C.; Sun, J.; Smith, C.; Wang, Y. Appl. Catal. A Gen. 2013, 467 (0), 91–97.

(11) Jørgensen, B.; Egholm Christiansen, S.; Dahl Thomsen, M. L.; Christensen, C. H.

J. Catal. 2007, 251 (2), 332–337.

(12) Carotenuto, G.; Tesser, R.; Di Serio, M.; Santacesaria, E. Catal. Today 2013, 203

(0), 202–210.

(13) Birky, T. W.; Kozlowski, J. T.; Davis, R. J. J. Catal. 2013, 298 (0), 130–137.

(14) Tsuchida, T.; Sakuma, S.; Takeguchi, T.; Ueda, W. Ind. Eng. Chem. Res. 2006, 45

(25), 8634–8642.

(15) Lebedev, S. V. Zh Obs. Khim 1933, No. 3, 698–717.

(16) Jones, M. D. Chem. Cent. J. 2014, 8 (1), 53.

(17) Chang, F.-W.; Yang, H.-C.; Roselin, L. S.; Kuo, W.-Y. Appl. Catal. A Gen. 2006,

304 (0), 30–39.

(18) Makshina, E. V.; Janssens, W.; Sels, B. F.; Jacobs, P. A. Catal. Today 2012, 198

(1), 338–344.

(19) De Baerdemaeker, T.; Feyen, M.; Müller, U.; Yilmaz, B.; Xiao, F.-S.; Zhang, W.;

Yokoi, T.; Bao, X.; Gies, H.; De Vos, D. E. ACS Catal. 2015, 3393–3397.

(20) Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A.

ChemSusChem 2014, 7 (9), 2505–2515.

(21) Ohnishi, R.; Akimoto, T.; Tanabe, K. J. Chem. Soc. Chem. Commun. 1985, No.

22, 1613–1614.

(22) Sekiguchi, Y.; Akiyama, S.; Urakawa, W.; Koyama, T.; Miyaji, A.; Motokura, K.;

Baba, T. Catal. Commun. 2015, 68, 20–24.

(23) Larina, O.; Kyriienko, P.; Soloviev, S. Catal. Letters 2015, 1–7.

(24) Sushkevich, V. L.; Ivanova, I. I.; Ordomsky, V. V; Taarning, E. ChemSusChem

2014, 7 (9), 2527–2536.

(25) Janssens, W.; Makshina, E. V.; Vanelderen, P.; De Clippel, F.; Houthoofd, K.;

Kerkhofs, S.; Martens, J. A.; Jacobs, P. A.; Sels, B. F. ChemSusChem 2014, 8 (6),

994–1008.

(26) Kvisle, S.; Aguero, A.; Sneeden, R. P. A. Appl. Catal. 1988, 43 (1), 117–131.

Page 121: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

101

(27) Kitayama, Y.; Satoh, M.; Kodama, T. Catal. Letters 1996, 36 (1–2), 95–97.

(28) Kitayama, Y.; Michishita, A. J. Chem. Soc., Chem. Commun. 1981, No. 9, 401–

402.

(29) Niiyama, H.; Morii, S.; Echigoya, E. Bull. Chem. Soc. Jpn. 1972, 45 (3), 655–659.

(30) Bhattacharyya, S. K.; Sanyal, S. K. J. Catal. 1967, 7 (2), 152–158.

(31) Bhattachariya, S.; Gangooli, N. D.; Avastkhi, B. N. J. Catal. 1965, 4 (6), 736.

(32) Bhattacharyya, S. K.; Avasthi, B. N. Ind. Eng. Chem. Process Des. Dev. 1963, 2

(1), 45–51.

(33) Sushkevich, V. L.; Ivanova, I. I.; Taarning, E. Green Chem. 2015.

(34) Baylon, R. a. L.; Sun, J.; Wang, Y. Catal. Today 2015, 1–7.

(35) Bhattacharyya, S. K.; Ganguly, N. D. J. Appl. Chem. 1962, 12 (3), 97–104.

(36) Jones, M. D.; Keir, C. G.; Iulio, C. Di; Robertson, R. A. M.; Williams, C. V;

Apperley, D. C. Catal. Sci. Technol. 2011, 1 (2), 267–272.

(37) Quattlebaum, W. M.; Toussaint, W. J.; Dunn, J. T. J. Am. Chem. Soc. 1947, 69,

593–599.

(38) Toussaint, W. J.; Dunn, J. T.; Jachson, D. R. Ind. Eng. Chem. 1947, 39 (2), 120–

125.

(39) Ostromislenskiy, I. J. Russ. Phys. Chem. Soc. 1915, No. 47, 1472 –1506.

(40) Gruver, V.; Sun, A.; Fripiat, J. J. Catal. Letters 1995, 34 (3–4), 359–364.

(41) Chieregato, A.; Velasquez Ochoa, J.; Bandinelli, C.; Fornasari, G.; Cavani, F.;

Mella, M. ChemSusChem 2014, 8 (2), 377–388.

(42) Zhang, M.; Gao, M.; Chen, J.; Yu, Y. RSC Adv. 2015, 5 (33), 25959–25966.

(43) Chizallet, C.; Costentin, G.; Che, M.; Delbecq, F.; Sautet, P.; Surface, D.; Marie,

P. J. Am. Chem. Soc. 2007, No. 19, 6442–6452.

(44) Shalabi, A. S.; El-Mahdy, A. M. Phys. Lett. A 2001, 281 (2–3), 176–186.

(45) Chizallet, C.; Bailly, M. L.; Costentin, G.; Lauron-Pernot, H.; Krafft, J. M.; Bazin,

P.; Saussey, J.; Che, M. Catal. Today 2006, 116 (2), 196–205.

(46) Di Valentin, C.; Del Vitto, A.; Pacchioni, G.; Abbet, S.; Wörz, A. S.; Judai, K.;

Heiz, U. J. Phys. Chem. B 2002, 106 (46), 11961–11969.

(47) Branda, M. M.; Rodríguez, A. H.; Belelli, P. G.; Castellani, N. J. Surf. Sci. 2009,

603 (8), 1093–1098.

(48) Kantorovich, L. N.; Holender, J. M.; Gillan, M. J. Surf. Sci. 1995, 343 (3), 221–

239.

(49) Sushko, P. V.; Shluger, A. L.; Catlow, C. R. A. Surf. Sci. 2000, 450 (3), 153–170.

(50) Chizallet, C.; Petitjean, H.; Costentin, G.; Lauron-Pernot, H.; Maquet, J.;

Bonhomme, C.; Che, M. J. Catal. 2009, 268 (1), 175–179.

(51) Aramendia, M. a.; Borau, V.; Jimenez, C.; Marinas, J. M.; Porras, A.; Urbano, F. J.

J. Mater. Chem. 1996, 6 (12), 1943.

(52) Bailly, M.; Chizallet, C.; Costentin, G.; Krafft, J.; Lauronpernot, H.; Che, M. J.

Catal. 2005, 235 (2), 413–422.

(53) Cornu, D.; Guesmi, H.; Krafft, J. M.; Lauron-Pernot, H. J. Phys. Chem. C 2012,

116 (11), 6645–6654.

(54) Corson, B. B.; Jones, H. E.; Welling, C. E.; Hinckley, J. A.; Stahly, E. E. Ind. Eng.

Chem. 1950, 42 (2), 359–373.

(55) Zhang, M.; Yu, Y. Ind. Eng. Chem. Res. 2013, 52 (28), 9505–9514.

(56) Redina, E. A.; Greish, A. A.; Mishin, I. V; Kapustin, G. I.; Tkachenko, O. P.;

Page 122: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

102

Kirichenko, O. A.; Kustov, L. M. Catal. Today 2015, 241, Part, 246–254.

(57) Fessenden, R. J.; Fessenden, J. S. In Organic Chemistry; Dodson, K., Ed.;

Brooks/Cole Publishing Company, 1995; pp 713–748.

(58) Harmony, M. D.; Laurie, V. W.; Kuczkowski, R. L.; Schwendeman, R. H.;

Ramsay, D. A.; Lovas, F. J.; Lafferty, W. J.; Maki, A. G. J. Phys. Chem. Ref. Data

1979, 8 (3), 619.

(59) Yang, Y.-J.; Teng, B.-T.; Liu, Y.; Wen, X.-D. Appl. Surf. Sci. 2015, 357, Part,

369–375.

(60) Chen, H.-Y. T.; Giordano, L.; Pacchioni, G. J. Phys. Chem. C 2013, 117 (20),

10623–10629.

(61) Cavalleri, M.; Pelmenschikov, A.; Morosi, G.; Gamba, A.; Coluccia, S.; Martra, G.

In Oxide-based Systems at the Crossroads of ChemistrySecond International

Workshop October 8-11, 2000, Como, Italy; A. Gamba, C. C. and S. C. B. T.-S. in

S. S. and C., Ed.; Elsevier, 2001; Vol. Volume 140, pp 131–139.

(62) Barborini, M.; Guidoni, L. J. Chem. Phys. 2012, 137 (22).

(63) Bučko, T.; Benco, L.; Hafner, J.; Ángyán, J. G. J. Catal. 2011, 279 (1), 220–228.

(64) Fessenden, R. J.; Fessenden, J. S. In Organic Chemistry; Dodson, K., Ed.;

Brooks/Cole Publishing Company, 1995; pp 275–308.

(65) Musolino, V.; Selloni, A.; Car, R. J. Chem. Phys. 1998, 108 (12), 5044.

(66) Pacchioni, G. J. Chem. Phys. 1996, 104 (18), 7329.

(67) Zhukovskii, Y. F.; Kotomin, E. a.; Borstel, G. Vacuum 2004, 74 (2), 235–240.

(68) Matveev, A.; Neyman, K.; Yudanov, I.; Rösch, N. Surf. Sci. 1999, 426 (1), 123–

139.

(69) Rodriguez, J. a.; Jirsak, T.; Chaturvedi, S. J. Chem. Phys. 1999, 111 (17), 8077.

(70) Rodriguez, J. a.; Jirsak, T.; Freitag, A.; Larese, J. Z. J. Phys. Chem. B 2000, 2

(100), 7439–7448.

(71) Angelici, C.; Meirer, F.; van der Eerden, A. M. J.; Schaink, H. L.; Goryachev, A.;

Hofmann, J. P.; Hensen, E. J. M.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ACS

Catal. 2015, 5 (10), 6005–6015.

(72) Boronat, M.; Corma, A.; Renz, M. J. Phys. Chem. B 2006, 110 (42), 21168–21174.

(73) Pasini, T.; Lolli, A.; Albonetti, S.; Cavani, F.; Mella, M. J. Catal. 2014, 317, 206–

219.

(74) Assary, R. S.; Curtiss, L. A.; Dumesic, J. A. ACS Catal. 2013, 3 (12), 2694–2704.

(75) Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A. Catal.

Sci. Technol. 2015.

(76) Arundale, E.; Mikeska, L. A. Chem. Rev. 1952, 51 (3), 505–555.

(77) Yamabe, S.; Fukuda, T.; Yamazaki, S. Beilstein J. Org. Chem. 2013, 9, 476–485.

(78) Miller, H. Conversion of a mixture of acetaldehyde and ethylene to butadiene.

US2377025 A, 1945.

(79) Isagulyants, V. I.; Safarov, M. G. Pet. Chem. U.S.S.R. 1966, 5 (3), 203–208.

(80) Wang, Y.; Wang, F.; Song, Q.; Xin, Q.; Xu, S.; Xu, J. J. Am. Chem. Soc. 2013,

135 (4), 1506–1515.

(81) Towns, J.; Cockerill, T.; Dahan, M.; Foster, I.; Gaither, K.; Grimshaw, A.;

Hazlewood, V.; Lathrop, S.; Lifka, D.; Peterson, G. D.; Roskies, R.; Scott, J. R.;

Wilkins-Diehr, N. Comput. Sci. Eng. 2014, 16 (5).

Page 123: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

103

Chapter 4

Surface Reaction Mechanisms Study of

MgO/SiO2 for Lebedev Process

Abstract .......................................................................................................................... 103

1. Introduction ....................................................................................................... 104

2. Results and Discussion ...................................................................................... 109

2.1. Catalyst activity and selectivity testing ....................................................... 109

2.2. In-situ DRIFT spectroscopy of MgO based catalyst surface hydroxyl

groups................................................................................................................... 110

2.3. Acid-base characterization of WK (1:1) catalyst using CO2 and pyridine

as probe molecules .............................................................................................. 113

2.4. In-situ DRIFT spectroscopy to monitor hydroxyl group reactivity during

the ethanol, acetaldehyde, crotonaldehyde and crotyl alcohol adsorption and

subsequent reaction on a WK (1:1) catalyst surface ........................................... 115

2.5. In-situ DRIFT spectroscopy of C2 (ethanol, acetaldehyde) and C4

(crotonaldehyde and crotyl alcohol) adsorption and reaction on WK (1:1)

catalyst surface as a function of temperature ...................................................... 119

2.5.1. C2 reactants and intermediates ............................................................... 119

2.5.2. C4 intermediates ..................................................................................... 127

2.6. DFT calculations ethanol, acetaldehyde, crotonaldehyde and crotyl alcohol

vibrational frequencies ........................................................................................ 130

2.7. In-situ DRIFT spectra for the ethanol, acetaldehyde, crotonaldehyde and

crotyl alcohol reaction on a WK (1:1) catalyst surface: the effect of the vapor

phase presence .................................................................................................... 136

3. Conclusions ........................................................................................................ 145

Supporting Information ............................................................................................... 148

References ...................................................................................................................... 151

Abstract

1,3-butadiene is an important commodity chemical and new, selective routes of

catalytic synthesis using green feedstock, such as ethanol, is of interest. For this purpose,

surface chemistry of MgO/SiO2 catalyst synthesized using wet-kneading was explored

during the reaction of ethanol and the corresponding reactive intermediates, including

Page 124: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

104

acetaldehyde, crotonaldehyde, crotyl alcohol using temperature programmed in situ

DRIFT spectroscopy combined with DFT calculations. Ethanol adsorption yielded several

physisorbed and chemisorbed surface species. Acetaldehyde exhibited high reactivity to

form crotonaldehyde. However, aldehyde intermediates resulted in strongly bound surface

species stable even at high temperatures, assigned to surface acetate, and/or 2,4-hexadienal

or polymerized acetaldehyde. Crotonaldehyde was reduced by ethanol to yield crotyl

alcohol via MPV mechanism. Crotyl alcohol, on the other hand, showed to be very reactive

and yield two different species on the surface, namely physisorbed and deprotonated that

would further desorb as 1,3-BD. Presence of gas phase hydrogen containing molecules,

such as ethanol, proved to be key in several reactive steps, including acetaldehyde

condensation step and crotonaldehyde reduction. Altogether, these data suggested a

complex reactive interactions between the surface hydroxyl groups, gaseous reactants and

surface bound reactive intermediates during the 1,3-BD formation. Future work is needed

to correlate vapor phase product evolution with the transient reactive surface intermediates

to examine trends leading to higher overall 1,3-BD selectivity.

1. Introduction

1,3-butadiene (1,3-BD) is an important commodity chemical with widespread

applications in polymer synthesis and as an organic chemistry intermediate.1 It is

commonly produced via crude oil cracking. 1,3-BD production can be affected by market

instability triggered by oil price fluctuations. This difficulty is compounded by the

emergence of shale gas, which suggests a need for an alternative and more sustainable

production method.2 For this reason, there has been a renewed interest in utilizing ethanol

Page 125: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

105

as a feedstock for 1,3-BD synthesis. Production from ethanol was used during World War

II by the USA and the USSR with a two-step process and a one-step catalytic process,

respectively, as demonstrated by Ostromislensky and Lebedev.3,4 Several reports since then

have highlighted the economic viability of the overall synthesis process with the one-step

catalytic process recently becoming a focal point.1,2,5 This single step process originally

utilized MgO/SiO2 catalyst with a 30-40% yield. Ethanol dehydrogenation to yield

acetaldehyde was identified as the rate-determining step in the generally accepted complex

reaction mechanism,6 with a large body of experimental work performed for elucidating

the reaction pathways.7–12 However, there is only a limited number of studies that focus

on the adsorbed reactive surface intermediates on MgO catalysts that utilize in-situ

spectroscopy to characterize the reaction intermediates under operating conditions.8,10,13,14

A one-step catalytic mechanism, as summarized in Figure 4.1, involves

dehydrogenation of ethanol to acetaldehyde on basic MgO sites1,10,15, followed by C-C

I II II

IV V VI

Figure 4.1. Main reaction mechanism proposed for ethanol to 1,3-butadiene via

Toussaint’s aldol condensation.

Page 126: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

106

coupling via the aldol condensation mechanism to yield crotonaldehyde.6,16,17

Crotonaldehyde can be further hydrogenated via MPV (Meerwein–Ponndorf–Verley)

reduction, either with ethanol or molecular H2 and the resulting crotyl alcohol dehydrated

to give 1,3-BD.16,18 Three computational studies by Chieregato et al. 10, Zhang et al.11 and

Taifan et al.17 attempted to unravel the reaction mechanism and the structures of the

reactive surface intermediates. Zhang et al. performed calculations using periodic density,

functional generalized, gradient approximation (GGA), with a focus on the very first step

of the overall reaction mechanism, the dehydrogenation of ethanol. A stepped MgO

surface was predicted to have a lower energy barrier than a flat surface for this reaction.11

The dissociation of ethanol on that surface was studied for three different surface sites, i.e.,

Mg5CO5C, Mg4CO4C, and Mg3CO3C and on the stepped surface, i.e. Mg4CO4C was shown to

have the lowest potential energy barrier for this reaction. Chieregato et al., on the other

hand, proposed an entirely different mechanism based on cluster type calculations and a

Gaussian basis set. They ruled out crotonaldehyde and crotyl alcohol as possible

intermediates and concluded that acetaldehyde would react with a carbanion, which

resulted from ethanol C-H bond cleavage.10 Taifan et al., for the first time, outlined a

complete reactive pathway for ethanol conversion to 1,3-BD by using periodic GGA

calculations. The pathways explored included an alternative Fripiat’s Prins mechanism and

a mechanism based on Ostromislensky’s hemiacetal rearrangement. They showed ethanol

dehydration to have an energetic barrier comparable to that of dehydrogenation. The

dehydration proceeded on Mg3c with surface O5c responsible for the initial proton transfer

and the resulting low coordination O2c and O4c hydroxyl group formation.

Dehydrogenation, on the other hand, took place in the vicinity of Mg3c and step O4c.

Page 127: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

107

Acetaldol (3-hydroxybutanal) formation proceeded via a 16 kcal/mol free energy barrier,

as calculated at 450 oC using harmonic/rigid rotor approximation. That took place when

acetaldehyde molecules were coordinated to Mg3c, Mg4c and Mg5c surface sites. Acetaldol

was not identified as a stable intermediate on a PES surface, as it immediately deprotonated

on reactive O4c sites. While computational modelling utilized low saturation Mg3c and O4c

sites for the overall reaction cycle, spectroscopic identification of the corresponding sites

and the reactive surface intermediates was less utilized. In particular, an adsorbed ethoxy

group on MgO was identified at 1119-1132 cm-1 in the temperature regime between 200

and 400 oC, when adsorbed ethanol was heated suggesting ethanol chemisorption.10 New

bands appeared at 1718 and 1143 cm-1 already at 150 oC and these bands were attributed

to acetaldehyde and new C-O containing surface species assigned to carbanion,

respectively. A transient peak at 1653 cm-1 together with one at 2957 cm-1 were assigned

to acyl or acetyl species. The reactive adsorbed intermediate at 1620 cm-1 was observed

and assigned to crotyl alcohol.19 C=O and C=C stretching vibrations at 1672 and 1649 cm-

1 observed above 350 oC were assigned to other C4 products, such as 1,3-BD,

crotonaldehyde and butanol. No adsorbed crotonaldehyde or acetaldol intermediates (the

latter one in agreement with Taifan et al.)17 were observed at lower (>250 oC) temperatures

but 1,3-BD formation was identified, suggesting that the aldol condensation mechanism

was not a key. In-situ DRIFTS was also the preferred technique used by Davis’ group13,14

and Ordomsky et al.8 to monitor the surface species during the reaction to n-butanol and

1,3-butadiene, respectively. Ethanol strongly adsorbed on MgO as both dissociated

ethoxide and a molecularly adsorbed ethanol. Dissociated ethanol exhibited a major peak

at 1132-1119 cm-1, while molecularly adsorbed ethanol was detected at 1058 cm-1. At

Page 128: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

108

higher temperatures, no aldol condensation was detected during the experiment, possibly

due to the very low conversion which was also supported by a small acetaldehyde formed

at 1711 cm-1.13,14 Aldol condensation of acetaldehyde was studied over MgO/SiO2 catalyst

and was suggested to instantaneously take place on the surface once acetaldehyde was

introduced to the IR cell, as shown by the band at 1643 cm-1, attributed to C=C stretch of

crotonaldehyde. Other than aldol condensation, acetaldehyde undergoes other side

reactions, namely condensation on the basic sites, as well as aldol condensation with

crotonaldehyde yielding 2,4-hexadienal, an unsaturated aldehyde.8

In this work we report a detailed study of wet-kneaded MgO/SiO2 catalyst surface

reactive sites and reactive intermediates during the ethanol conversion to 1,3-BD. Where

necessary, data are also reported for pure MgO model catalyst. Wet kneading (WK) of

MgO/SiO2 catalyst has been shown to result in an active catalysis towards 1,3-BD

formation from ethanol.1,15,20–23 In this study, we prepared calcined MgO and MgO/SiO2

catalysts using a wet kneading method with an MgO:SiO2 mass ratio of 1:1. We used in

situ Diffuse Reflectance Infrared Spectroscopy (DRIFTS) and the corresponding proposed

reactants and proposed reactive intermediates, including ethanol, acetaldehyde, crotyl

alcohol, crotonaldehyde, as shown in Figure 4.1. To aid in the observed peak assignment,

quantum chemical calculations using periodic boundary conditions and PBE functional

were used.

Page 129: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

109

2. Results and Discussion

2.1. Catalyst activity and selectivity testing

Figure 4.2 depicts the catalytic activity testing of the synthesized wet-kneaded

MgO/SiO2 catalyst, WK (1:1). The experiment were carried out at 723 K (450°C) at several

WHSV (hr-1) ranging from 0.78 to ~2 hr-1. WHSV plays a very important role in

determining the catalyst’s activity, since it represents the catalyst-to-reactant ratio. The

conversion decreased with increasing WHSV from ~87% to ~60%. Selectivities of selected

products, i.e. acetaldehyde, ethylene, and 1,3-butadiene, were relatively unaffected by

WHSV. At very high conversion, there were several other byproducts, such as butenes,

propene, ethers and some aromatic compound that coked the catalyst; this led to carbon

balance of 60-80%. High 1,3-butadiene selectivity was achieved with this catalyst, 35-

40%, without the addition of routinely used transition metal oxide promoter.1,2 Similar

conversion-yield values were previously reported by Weckhuysen’s group, where different

methods of preparation and precursors were explored to find the best working catalyst.21

Table 4.1 shows comparison between catalyst in this work and previously used wet-

kneaded MgO/SiO2 catalysts.

Table 4.1. Catalytic activity comparison of WK (1:1) with previously investigated wet-

kneaded synthesized catalysts.

Catalyst T (K) WHSV (h-1) XEtOH

(%)

YBD

(%)

PBD

(gBD g-1cat h-1) Ref

WK (1:1) 723 1.1 ~84 33 0.4 This

work

WK-a 623 0.15 50 42 0.06 Makshina

et al.23

WK-b 698 1.1 ~67 35 0.25 Angelici

et al.21

Page 130: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

110

2.2. In-situ DRIFT spectroscopy of MgO based catalyst surface hydroxyl groups.

We begin by investigating the hydroxyl groups present on a dehydrated WK (1:1)

surface by comparing them with MgO and SiO2. In-situ IR spectroscopy results for the

pure MgO surface shown in Figure 4.3 were obtained after heating (dehydrating) the

sample at 773 K, typical for the ethanol catalytic reaction to form 1,3-BD, in air and cooling

down to 373 K temperature. Spectra show two high basicity (low coordination) peaks in

the hydroxyl region at 3765 and 3745 cm-1, while several broad peaks are also present at

3700-3400 cm-1, namely, 3660, 3547 and 3465 cm-1. The higher stretching frequency is

related to a more isolated (and basic) hydroxyl group, while the lower one is often assigned

to multi-coordinated hydrogen bonded hydroxyls.27,39–43 In general, there have been six

structural hydroxyl group models proposed to exist on MgO. Anderson et al. proposed two

kinds of hydroxyl groups on the MgO surface: hydrogen bond acceptor and hydrogen bond

donor.41 Their model was subsequently refined by Shido et al., where the two regions

Figure 4.2. Conversion (●) and selectivity of main products (■ acetaldehyde; ▲

ethylene; ♦ 1,3-butadiene) at different WHSV. Reaction conditions: T=723 K, Qtot = 50

cm3/min, Mcat=0.2 g, P0EtOH = 2.72; 3.77; 5.15; 6.96 kPa.

Page 131: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

111

could be classified in further detail based on the coordination numbers of the Mg and O

atoms.42 Coluccia, Morrow, and Knozinger each proposed three different models with one

uniting characteristic: the inclusion of an isolated hydroxyl group as the sharp band at the

high wavenumber region and hydrogen bond donor - multicoordinated hydroxyl groups in

the lower wavenumber region.39,40,43 Most recent models combined DFT and infrared

spectroscopy studies to show that the most isolated single coordinated (O1C-H) group does

not exist: it immediately transforms into O3C-H and O4C-H at a lower temperature and into

O2C-H at an elevated desorption temperature, and thus a new model was proposed.44 The

bands observed at 3765 and 3745 cm-1 in our work agree well with those reported in the

literature. Those bands have been assigned to low coordinated O1c-H or O2c-H hydrogen

bond acceptors or O4c-H and O5c-H coordinated isolated groups on valleys and edges of

the MgO crystallites.39,44 The peaks below 3650 cm-1 are in general attributed to multi-

coordinated hydrogen bond donor hydroxyl groups44, thus presenting a rather complex

picture of the reactive MgO surface.

Figure 4.3. In-situ DRIFTS spectra acquired of dehydrated (temperature programmed

to 773 K at 10 oC/min under air and cooled down to 100 oC) MgO, MgO WK (1:1)

catalysts and SiO2. Only hydroxyl region of 3800 to 3200 cm-1 is shown. Spectra are

acquired at 100 °C.

Page 132: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

112

The WK (1:1), on the other hand, exhibited four major peaks at 3745, 3725, 3705,

and 3680 cm-1. For comparison, dehydrated spectra of calcined SiO2 are also shown. There

are two bands present for SiO2, a sharp one at 3745 cm-1 and a broad band at 3700-3450

cm-1. The sharp peak is typical for the isolated silanol (Si-OH) vibration with a small

contribution from the geminal silanol group (HO-Si-OH), while the broad band is formed

from the contribution of the hydrogen bonded vicinal silanol groups.45 The 3745 cm-1 peak

is also observed in WK (1:1), albeit at the lower intensity, which suggests that isolated

silanol groups are consumed during the wet-kneading interaction with MgO. The other

three peaks present, 3725, 3705, and 3680 cm-1, are unique to the WK (1:1) structure. The

latter peak has previously been assigned to magnesium silicates, due to its formation in the

presence of silica.7,19,46 It has previously been observed as a mineral lizardite hydroxyl

group at 3686 cm-1.47 The relatively low FWHM (Full Width at Half Maximum) of the

peak suggests that this group might be isolated, rather than hydrogen bonded, consistent

with the crystalline structure of the lizardite.48 Peaks at 3725 and 3705 cm-1 are more

difficult to assign directly, since none of the magnesium silicate compounds exhibit

hydroxyl stretches above 3700 cm-1.47 It can be proposed that the interaction of MgO and

SiO2 during wet kneading increases the formation of hydroxyl groups that are already

present on MgO itself, i.e., wet-kneading results in more defects that produce the said

hydroxyl group or that peaks could originate from Mg-OH interacting with nearby SiO2

surface sites. The peak at 3725 cm-1 is rather intriguing due to the fact that it was not

observed by other groups. We tentatively assign the peaks at 3725 and 3705 cm-1 to the

isolated O4c-H and O5c-H coordinated groups formed in the presence of the amorphous

SiO2 (SiMg4cO4c and SiMg4cO5c). This is also consistent with the decrease in intensity of

Page 133: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

113

the 3765 and 3745 cm-1 hydroxyl groups present on MgO but not on WK (1:1), where

incorporation of amorphous Si-O-Mg linkages could result in the frequency shift towards

lower wavenumbers.

2.3. Acid-base characterization of WK (1:1) catalyst using CO2 and pyridine as

probe molecules

Characterization of the basic sites present on WK (1:1) was carried out by

performing in-situ DRIFTS using CO2 as a probe molecule. Figure 4.4a depicts the spectra

of adsorbed CO2 species at different temperatures. The basicity was previously reported to

originate from MgO, and with no contribution from SiO2.2 There are three broad, main

peaks present on the spectra, located at 1650, 1531, and 1405 cm-1. Judging from the

carbonate υ3 frequency split, the last two peaks originate from monodentate carbonate,

assigned to υ3 as and υ3 s, respectively.49 The peak at 1650 cm-1 could originate from either

bidentate carbonate or monodentate bicarbonate. However, bicarbonate would exhibit a

peak at around ~1250-1200 cm-1, which is non-existent in this case. Bidentate carbonate

assignment is more likely than bicarbonate, given the broad peaks exhibited in this spectra,

where the accompanying υ3 s would be convoluted as a shoulder to the peak at 1405 cm-1.

Furthermore, the basic site strength can be determined by the surface species present.

Monodentate carbonate is typically more stable than bidentate carbonate, while bicarbonate

is the least stable.2,49 Hence, the strong basic sites are assigned to monodentate carbonate,

while medium-strength and weak basic sites are assigned to bidentate carbonate and

bicarbonate, respectively. From the spectra, the bidentate carbonate is far less intense than

monodentate carbonate, indicating the more basic nature of the MgO/SiO2 WK (1:1).

Several other methods of preparation, such as sol-gel19 and incipient wetness

Page 134: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

114

impregnation2, yield catalysts with limited amount of strong basic sites. However, it should

be noted that at the reaction temperature, ~673-723 K, the CO2 species in our catalyst are

mostly absent. This indicates that the basic sites present on the catalyst gradually lose

strength at elevated temperature.

The acidity of the catalyst was characterized using pyridine as the probe molecule

(Figure 4.4b). Peaks at 1445 and 1605 cm-1 indicates the presence of strong Lewis acid

sites, while the peak at 1577 cm-1 is for weak Lewis acid sites.50 Peak at 1490 cm-1 is

assigned to a combination band of Lewis and Brønsted acid sites, while Brønsted acid site

itself should exhibit a peak at 1540 cm-1, which is not present on our catalyst. The peak at

1595 cm-1 does not represent any acid sites, instead, it was assigned to hydrogen-bound

pyridine.50 The absence of Brønsted acid sites were also observed by previous

investigators, given the basic nature of the catalyst.2,7,19,51 However, the intensity of the

strong Lewis acid sites is a dominant feature on this spectra indicating that the catalyst

possess a significant amount of strong Lewis acid sites, relative to the weaker Lewis acid

sites. SiO2 by itself is known to be slightly acidic, contributing to the weak Lewis acid

(a) (b)

Figure 4.4. DRIFTS spectra of adsorbed (a) CO2 and (b) pyridine on WK (1:1) catalyst

at different temperatures to probe the catalyst’s basicity and acidity at relevant

temperatures.

Page 135: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

115

sites, while the rest of the Lewis acid sites are combination of defect sites of MgO and the

interaction between SiO2 and MgO.1,2,10,51

2.4. In-situ DRIFT spectroscopy to monitor hydroxyl group reactivity during the

ethanol, acetaldehyde, crotonaldehyde and crotyl alcohol adsorption and subsequent

reaction on a WK (1:1) catalyst surface.

Spectra for those hydroxyl groups in the 3800 to 3200 cm-1 regions during WK

(1:1) reaction with ethanol, acetaldehyde, crotonaldehyde and crotyl alcohol are shown in

Figure 4.5 and tabulated in Table 2. Subtracted adsorbed molecule spectra as a function

of temperature are shown in black, while red dotted spectra are for 373 K hydroxyl groups

reacting upon vapor phase molecule adsorption with catalyst spectrum subtracted.

Notably, the catalyst sample surface was treated at 773 K beforehand; thus, the hydroxyl

groups observed in Figure 4.5 are transient reactive groups formed and released during the

organic molecule adsorption/reaction. Upon adsorption of organic molecules, negative

peaks appeared on all the assigned WK (1:1) hydroxyl groups, i.e. 3747, 3725, 3705, and

3680 cm-1. The adsorption behavior is very different for alcohols – ethanol and crotyl

alcohol – and for aldehydes – acetaldehyde and crotonaldehyde, as shown by the different

intensities of the negative peaks. The alcohols have less affinity to the peaks at 3705 and

3680 cm-1, while aldehydes have no preference on which hydroxyl group to coordinate.

Alcohols’ interactions with MgO surfaces include both molecular adsorption on native

hydroxyl groups52, as well as their displacement via chemisorption, which involves basic

site – Lewis acid site pairs, that will produce adsorbed water as the byproduct.19,53 Positive

hydroxyl peaks in the alcohol cases can indicate new hydroxyl vibrations, due to the newly

Page 136: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

116

formed groups via displacement, while the increased hydrogen bonding demonstrates that

some of the native hydroxyl groups only weakly-bind the molecular ethanol. Aldehyde

adsorption, on the other hand, typically takes place via two surface species: on a surface

hydroxyl group via an unstable, protonated intermediate and on a lone pair of oxygen atoms

as a more stable species, typically indicated by the red-shifted C=O stretching vibration at

1650-1680 cm-1.54 Figure 4.5 shows that acetaldehyde adsorbs differently from

crotonaldehyde, that the peak at 3725 cm-1 is not significantly consumed, as compared to

that of crotonaldehyde. We assume that this is due to crotonaldehyde’s π-electron cloud,

which makes the molecule more activated toward consuming the hydroxyl group related

to the 3725 cm-1 peak. For all experiments, the adsorption results in the positive peak at

~3684 cm-1, indicating different hydroxyl group coordination, or a more intense hydrogen

bonding. This positive peak is more intense for acetaldehyde and crotonaldehyde, possibly

due to the formation of new alcoholic species, rather than in the case of alcohols, which

are simply hydroxyl groups displacement.

Table 4.2. Surface hydroxyl group vibrational frequencies during ethanol, acetaldehyde,

crotonaldehyde and crotyl alcohol adsorption on WK (1:1).

Experimental (cm-1)

Ethanol Acetaldehyde Crotonaldehyde Crotyl

alcohol

Assignment

ν

(Mg-

OH)

3748

(3761)

3725

(3721)

-

3680

3755 (3757)

3721

3710

3680

3740

3721

3710

3680

3751

3723

-

3678

Mg4cO4c

SiMg4cO4c

SiMg4cO5c

Mg3Si3O5(OH)4

According to Figure 4.3, the peak at 3747 cm-1 is a combination of both a silica

isolated silanol peak and the basic MgO hydroxyl group. As the temperature is increased,

the former sharply loses intensity, while the latter slowly gains intensity. This trend is true

Page 137: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

117

for all the intermediates adsorbed on the surface. Furthermore, this MgO peak splits at a

higher temperature, indicating the presence of a second peak, at lower wavenumber, which

translates to higher coordination. This further splitting was previously observed by

Knözinger et al.39 All other surface hydroxyl groups undergo a significant decrease in

intensity, while also being accompanied by the emergence of their shoulders at a lower

wavenumber as the temperature is increased. One intriguing observation is that those

neighboring hydroxyl peaks are all red-shifted from the native hydroxyl peaks. The thermal

effect on the surface seems to rearrange the hydroxyl group coordination to achieve more

thermodynamically stable configurations, i.e., there are no new hydroxyl groups being

formed.

Putting the rearrangement of the hydroxyl groups aside, increasing the temperature

also led to desorption of the surface species. The release of the hydroxyl groups can be

explained by the flattening shoulder at ~3684 cm-1. These hydroxyl groups were made

during the alcohol/acetaldehyde adsorption. However, native hydroxyl peaks that were

consumed during the initial adsorption keep decreasing in intensity as well. This

continuous decrease indicates that these peaks, in particular at 3747 and 3680 cm-1, are not

fully consumed during the adsorption, i.e. they are relatively less reactive. The remaining,

unconsumed hydroxyl groups of these types undergo further thermal change by achieving

thermodynamically more stable coordination, shown by the increase of the neighboring

hydroxyl peak. For the case of aldehydes, increasing the temperature would both convert

the unstable protonated intermediate into the more stable compound, which is coordinated

to Lewis acid sites.

Page 138: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

118

Figure 4.5. In-situ DRIFTS spectra in the hydroxyl group region of 3800 – 3200 cm-1

acquired of ethanol, acetaldehyde, crotonaldehyde and crotyl alcohol on WK (1:1)

catalyst. Sample vapor was adsorbed on the sample surface and temperature ramped

up from 373 to 723 K while spectra being recorded. In-situ DRIFTS dehydrated

catalyst spectrum at 100 °C was used as a reference.

Ethanol Acetaldehyde

Crotonaldehyde Crotyl alcohol

Page 139: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

119

2.5. In-situ DRIFT spectroscopy of C2 (ethanol, acetaldehyde) and C4

(crotonaldehyde and crotyl alcohol) adsorption and reaction on WK (1:1) catalyst

surface as a function of temperature.

2.5.1. C2 reactants and intermediates.

When ethanol was adsorbed on WK (1:1), several peaks in the C-H region were

observed between 2985 and 2907 cm-1, as well as at 2720 cm-1, as shown in Figure 4.6 and

tabulated in Table 4.3. The former peaks were attributed to the combination of (CH3)

and (CH2) vibrational modes with the corresponding bending modes located at 1454 and

1418 cm-1, while the assignment of the latter peak is not straightforward. While generally

within the (CH3) spectral region, it can’t be unambiguously assigned. As later shown by

DFT calculations, that peak can be assigned to a frustrated (OH) mode of dissociated

(chemisorbed) ethanol on a Mg4c site. Peaks at 1380 and 1338 cm-1 can be assigned to the

wagging modes of CH2 and CH3. Interestingly, a peak at 1624 cm-1 was observed at 373

K together with a negative peak at ~1670 cm-1 after physisorbed ethanol adsorption. The

peak at 1670 cm-1 is the native bidentate carbonate asymmetric ν3 vibration that persisted

during the sample treatment, which we propose to be displaced due to the reactive ethanol

adsorption on the surface.49 This bidentate carbonate peak would typically be accompanied

by its asymmetric counterpart at ~1370 cm-1,49 however, this spectral region also shows

adsorbed ethanol vibrations. The splitting of the ν3 vibration is typically used to identify

the coordination of the carbonate, since the degree of symmetry lowering caused by surface

coordination is well-known to split the vibration differently.55 For instance, the

monodentate split is ~100 cm-1 (1415 split to ~1400 and ~1500 cm-1), bidentate split is

~300 cm-1, and bridged carbonate split is ≥ 400 cm-1.55 While surface carbonates are

Page 140: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

120

typically stable under low O2 and high CO2 concentration, they are typically unstable under

reaction conditions.56 DFT study also showed that ethanol adsorption is

thermodynamically more stable than that for CO2, with ED = 10.5 - 13.5 kcal/mol and 7.1

- 9.4 kcal/mol, respectively,17,57 hence suggesting that chemisorption of ethanol can

displace surface carbonate species. There are several possibilities in assigning the peak at

1624 cm-1. At this wavenumber region, hydrocarbons exhibit C=C stretches, as well as a

distorted C=O stretch. Stable bands at ~1580 cm-1 are also expected for the surface acetate,

even though an accompanying peak must be present around ~1400 cm-1 It is unlikely, that

acetate is formed at 373 K.58 One possible explanation for this is the formation of the

coadsorbed water, presumably formed by dissociation of ethanol that results in the

rearrangement of the hydroxy group when ethanol is adsorbed on the surface. The

formation of water on the surface was previously observed by Busca’s group on γ-Al2O3,

where vapor-phase water was formed right after ethanol was introduced to the surface.53

The corresponding ν(C-O) vibration at ~1100 cm-1 was not observed because of

the predominant vibrations of SiO2 in the spectral region of 1200 to 900 cm-1. However,

peaks at 1140, 1126, 1104 and 1058 cm-1 were observed at 373 K on pure MgO samples,

as shown in the Figure 4.6 inset. The peaks at 1140, 1104, and 1058 can be assigned to

the ethanol species on the surface, both chemisorbed surface ethoxide and physisorbed

ethanol, with spectral shifts observed due to the difference in interaction with and

adsorption on various accessible sites. The shifts between the physisorbed, chemisorbed,

and vapor-phase ethanol were also previously observed by Branda and Birky.14,30 Peaks at

1104 and 1058 cm-1 were assigned to δ(OH) and ν (C-O) of surface ethoxy, respectively,

as also observed by Davis and Cavani groups.10,13,14 The peak at 1140 cm-1 gained intensity

Page 141: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

121

after the cell was evacuated and heated to 476 K. This peak was previously assigned to ρw

(CCO) by Davis’ group with the peak located at 1122 cm-1 on MgO.14 Notably, the peak at

1126 cm-1 appears at a higher temperature, ~473 K, which indicates the formation of a new

species. This peak can be assigned to C-C stretch of the adsorbed acetaldehyde.59

Table 4.3. Vibrational frequencies and their assignments for ethanol, acetaldehyde,

crotonaldehyde and crotyl alcohol adsorption on WK (1:1)

Assignment

Experimental (cm-1)

Ethanol Acetaldehyde Enolate Crotonaldehyde Crotyl

alcohol

ν (OH) 2720 - - - -

ν (CH3) 2985 3037, 2967,

2935 - 2967, 2935

3017,

2965

ν (CH2) 2937,

2907, 2881 - - -

2949,

2840

Fermi CH3 2882, 2845

ν (CH) - 2743 - 3032, 2882,

2845

ν (C=O) - 1716, 1680,

1633 -

1716, 1680,

1670 -

ν (C=C) - - 1600,

1578 1600, 1574 1602

δ (CH2) 1454 - - - 1380

δ (CH3) 1418 - - 1456, 1434 1368

ρw (CH) 1380 - - - -

ρw (CH2) - - - - 1441

ρw (CH3) 1338 1456, 1434,

1382 - 1346 1456

ρw (CHO) - 1284 - 1382 -

As the temperature is increased to 723 K, the aforementioned bands start decreasing

in intensity, giving rise to several new bands, including those at 2958, 1653, 1604 and 1581

cm-1. The experiment was carried out under inert gas flow, which prevents chemistry

beyond dehydrogenation and dehydration from happening. The inert atmosphere

encouraged desorption to take place rather than surface reaction, further limiting the

Page 142: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

122

reaction to dehydration and dehydrogenation, which are elementary reactions of ethanol.

Experiments under constant reactant flow, where ethanol is constantly supplied to the

surface, were also done and discussed thoroughly in Section 3.7. The first two peaks

appeared intermittently from 473 K up to 623 K. The peak at 1653 cm-1 can be assigned to

the ν (C=O) of acetaldehyde59 adsorbed on an Mg3c surface site with the assignment

confirmed by our DFT calculation. This peak was also previously assigned to adsorbed

crotonaldehyde, since the crotonaldehyde C=O stretch frequency is also typically lowered

upon adsorption.8 The crotonaldehyde vapor-phase exhibits C=O stretch at ~1691 cm-1, as

shown in Table S4.1. The appearance of the former can then be related to acetaldehyde as

well, as the CH2 stretch was also observed by Raskó et al. at 2960 cm-1 on TiO2 and

Ordomsky et al. at 2950 cm-1 over SiO2, ZrO2/SiO2, and MgO/SiO2.8,54 Aldehyde presence

can typically be spotted by its unique carbonyl CH stretch at ~2700 cm-1, however this

peak does not exist in our spectra. The absence of this peak is due to the distorted

acetaldehyde -CHO group, where the C-O bond is now elongated, and the band will shift

to a higher wavenumber at ca. 2800 cm-1, as explained later by DFT results. Ordomsky et

al. had this same observation where the carbonyl CH stretch was only present on the spectra

when there is vapor-phase acetaldehyde on the surface.8 Upon increasing the temperature,

two peaks at 1604 and 1581 cm-1 start gaining intensity; these speaks can be assigned to a

C=C group. We have previously shown that acetaldehyde transformation to surface enolate

has very low activation energy on an Mg3c surface site, which becomes the basis of our

assignment of one of these C=C stretching bands to surface enolate.17 The gas-phase keto-

enol tautomerization of acetaldehyde thermodynamically favors the keto (aldehyde)

form.60 However, Palagin et al. spectroscopically observed proton transfer from

Page 143: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

123

acetaldehyde to SnBEA zeolite and attributed this to the surface enolate formation. Further

complicating the assignment of these C=C stretching bands is the rather low Gibbs free

energy barrier of ~16 kcal/mol17 for aldol condensation, which opens up the possibility for

crotonaldehyde formation.

Conflicting assignments have also previously been made for the C=C stretching

bands originating from ethanol adsorption on MgO, where they were also previously

assigned to surface acetate,52,58 and/or 2,4-hexadienal, aldol condensation product of

crotonaldehyde and acetaldehyde.8 The latter was actually shown to form on basic surface

sites at close to dry ice temperatures.61,62 The formation of surface acetate was previously

postulated to take place through an acyl intermediate, which yields a vibrational band at

~1690 cm-1.59 There are several reasons why acyl intermediate is highly unlikely to be

formed on the basic surface: the base-induced Cannizzaro reaction suggests that the

aldehyde needs to be lacking hydrogen in the α-position63, and hence, the only way

Figure 4.6. In-situ DRIFTS spectra acquired of ethanol on WK (1:1) catalyst in 3200

to 1000 cm-1 spectral region. Ethanol was adsorbed on the sample surface and

temperature increased from 373 to 723 K while spectra being recorded. In-situ DRIFTS

spectra of the sample surface with no adsorbate present at every corresponding

temperature were used for reference. In-situ DRIFTS spectra acquired for ethanol

adsorbed on MgO are shown in the inset for 1200 to 1000 cm-1 spectral region.

MgO

Page 144: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

124

acetaldehyde can form an acetate is from the strong Lewis acid-driven Tischenko

reaction.64 From the broad shape of the peak it can be concluded that surface acetaldehyde

undergoes enol transformation to yield surface enolates, which then either polymerize

and/or undergo aldol condensation to yield higher aldehydes and bulky aromatic structures.

Collectively, these data show that in addition to physisorption, ethanol chemisorbs

as two different surface species by displacing carbonate structure and producing water as

a byproduct of its adsorption. In addition to the dissociative chemisorption, ethanol adsorbs

as a semi-dissociated species, where the removed proton is still interacting strongly with

the ethoxide, as shown by the frustrated -OH vibration. Part of this structure further

undergo dehydrogenation, which is shown by the acetaldehyde C=O peak, and further

reacts to make C=C containing molecule(s).

Acetaldehyde is often cited as an important intermediate during the catalytic

ethanol transformation to 1,3-BD and has been proposed to have a major role in the overall

reactive mechanism.1,6,23 To further explore the relevant surface intermediates chemistry

and corroborate the assignment of the ethanol IR data, we carried out IR temperature

programmed desorption experiments with acetaldehyde as the probe molecule.

Acetaldehyde IR spectra typically exhibit peaks at 1716, 1680 and 1590 cm-1 which were

related to vapor-phase C=O, adsorbed C=O, and C=C stretches, respectively.65 Other

bands stemming from acetaldehyde adsorption on bifunctional metal oxide surfaces

previously observed were located in the 1450-1350 cm-1 region with conflicting

assignments to either surface acetates66 or, more conservatively, to bending modes of CH,

CH2, and CH3.65,67 Acetaldehyde adsorption was studied over MgO based catalysts, where

on Ni/MgO, it was claimed to undergo several reactions, such as a Cannizzaro reaction

Page 145: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

125

which yielded surface acetate peaks around ~1580 and ~1425 cm-1 68 aldol condensation,

as previously observed on MgO/SiO2 by IR spectroscopy.8

Acetaldehyde adsorption on WK (1:1) is shown in Figure 4.7 and tabulated in

Table 4.3. In the C-H region, several bands were observed at 3027, 2967, 2935, 2882,

2845, and 2743 cm-1 upon acetaldehyde introduction to the surface. The first three bands

can be directly assigned to the stretching mode of CH3, while CH3 Fermi resonances can

be assigned to 2882 and 2845 cm-1, These depend on the adsorption site.59 A possible

assignment can also be made that one belongs to acetaldehyde, while the other one to

crotonaldehyde. The weak band at 2743 cm-1 is the signature stretching mode of aldehyde

-CHO group with broadening due to possible crotonaldehyde formation at the low

temperature. The peaks at 1716, 1680, 1633, 1600 and 1578 cm-1 are all related to C=O

and C=C stretches, depending on the binding site and nature of the adsorption/perturbation

of the C=O group. Acetaldehyde undergoes aldol condensation17 and polymerization61,62

with very low energy barriers on basic surfaces, including MgO. The reactivity of

acetaldehyde on the basic surface is the main reason for its rapid deactivation when

acetaldehyde is present in large concentrations.8,69 Peaks at 1456, 1434, 1382, 1346 and

1284 cm-1 can be used to confirm the existence of an acetaldehyde molecule and its reaction

products, such as polymerized acetaldehyde and possible enolate isomerization-aldol

condensation products on the surface. The first two of those peaks, along with the one at

1346 cm-1 , are assigned to the wagging mode of CH3, the mode that polymerized aromatic

acetaldehyde is lacking.59 The peak at 1382 cm-1 and 1284 cm-1 can be assigned to the

bending mode of the acetaldehyde CH3 group59 and wagging mode of the acetaldehyde

CHO group, respectively. However, these bending modes can overlap with those

Page 146: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

126

originating from surface enolate, as a result of keto-enol tautomerization, and further

detailed assignments will be made on DFT calculations. There were no peaks observed

below 1300 cm-1 due to the strong SiO2 vibrations and pure MgO data was used to

investigate that spectral region, as shown in the Figure 4.7 inset. In this region, three bands

are observed at 1264, 1107, and 1066 cm-1 when acetaldehyde is observed on MgO. These

bands were previously assigned to η (C-O), η (C-C), and ν (C-C), respectively.59

Analysis of the IR data of both ethanol and acetaldehyde demonstrates that the

surface ethoxy species react to make adsorbed surface acetaldehyde, that, in turn, further

desorb, polymerize, couple, or isomerize as shown by the presence of the peaks at ~1600

and ~1580 cm-1, where intensity in the ethanol spectra is much less than that of

acetaldehyde. This lack of intensity is due to the fact that ethanol would also desorb from

the surface and the steep activation energy, i.e. 39.6 kcal/mol for the dehydrogenation

Figure 4.7. In-situ DRIFTS spectra acquired of acetaldehyde on WK (1:1) catalyst in

3200 to 1000 cm-1 spectral region. Acetaldehyde was adsorbed on the sample surface

and temperature increased from 373 to 723 K while spectra being recorded. In-situ

DRIFTS spectra of the sample surface with no adsorbate present at every corresponding

temperature were used for reference. In-situ DRIFTS spectra acquired for acetaldehyde

adsorbed on MgO are shown in the inset for 1200 to 1000 cm-1 spectral region.

MgO

Page 147: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

127

reaction.17 The acetaldehyde formed during the experiments has higher affinity to the basic

surface, which is confirmed by the intense CH stretching and CH3 wagging modes of the

molecule being on the surface even at 723 K. Ordomsky, et al. also acknowledged the

reactivity of the basic catalyst, which, in turn, results in the strong adsorption of

acetaldehyde and/or its higher self-reaction products on the surface.8

2.5.2. C4 intermediates.

DRIFT spectra of crotonaldehyde adsorbed on the WK (1:1) surface are shown in

Figure 4.8. Notably, peaks observed in the spectra are identical with the ones found for

acetaldehyde in Figure 4.7 and tabulated in Table 4.3, except for the relative intensities of

several peaks, such as 1716, 1680, 1650, 1456, and 1434 cm-1. The similarities between

the two spectra suggest potential overlaps between peaks from both aldehydes during the

Figure 4.8. In-situ DRIFTS spectra acquired of crotonaldehyde on WK (1:1) catalyst in

3200 to 1000 cm-1 spectral region. Crotonaldehyde was adsorbed on the sample surface

and temperature increased from 373 to 723 K while spectra being recorded. In-situ

DRIFTS spectra of the sample surface with no adsorbate present at every corresponding

temperature were used for reference.

Page 148: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

128

infrared analysis of the surface reactive intermediates. The almost identical spectra make

a solid case for the deduction of acetaldehyde spontaneous coupling to crotonaldehyde via

aldol condensation at 373 K followed by dehydration, which can also be confirmed by the

shoulder at ~1633 cm-1 in Figure 4.7. The interpretation of aldol condensation at 373 K

can be justified by our supplementary TPSR experiments (not shown), which showed that

crotonaldehyde (m/z=70) appeared right when the surface became saturated with

acetaldehyde. The spontaneous reaction of acetaldehyde aldol condensation to yield

hydroxy-butanal and crotonaldehyde had also been previously observed over HZSM-5,65

TiO2, CeO2, and Al2O3.54,58,59 However, infrared spectra alone are not capable of

decoupling both intermediates.

The adsorption of crotyl alcohol on the WK (1:1) surface, as shown in Figure 4.9

and tabulated in Table 4.3, provides more insight for the C=C vibration region. Our data

Figure 4.9. In-situ DRIFTS spectra acquired of crotyl alcohol on WK (1:1) catalyst in

3200 to 1000 cm-1 spectral region. Crotyl alcohol was adsorbed on the sample surface

and temperature increased from 376 to 723 K while spectra being recorded. In-situ

DRIFTS spectra of the sample surface with no adsorbate present at every corresponding

temperature were used for reference.

Page 149: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

129

show relatively few peaks and demonstrate the selective nature of the crotyl alcohol

transformation on this bifunctional catalyst. Peaks at 3017, 2965, 2949 and 2840 cm-1 were

assigned to the CH3 and CH2 stretches of adsorbed crotyl alcohol, while peaks at 2932,

2867 and 2738 cm-1 were assigned to both and also to vapor-phase crotyl alcohol.70

Bending and wagging modes of the CH groups can also be seen in the 1400-1300 cm-1

region, where peaks at 1456 and 1441 cm-1 were assigned to ρw (CH2) and ρw (CH3) and

1380 and 1368 cm-1 - to δ (CH2) and δ (CH3). A unique feature of the adsorption is shown

by the presence of an intense, sharp peak at 1602 cm-1 upon adsorption, which red-shifted

~70 cm-1 from the vapor-phase crotyl alcohol peak at 1675 cm-1.70 This peak is assigned to

the C=C stretch of adsorbed 1,3-BD. This was also observed by Wenig and Schrader during

their crotyl alcohol in-situ IR experiments over V-P-O catalyst.71 The accompanying water

peak at ~1650 cm-1 can be seen as a shoulder for the intense 1600 cm-1 peak. Our DFT

calculations discussed later also show that at this adsorbed state, there was no higher CH

peak observed at ~3100 cm-1, due to the elongated C=C group of the C4 structure.

Following a temperature increase to 473 K, a shoulder appeared at 1620 cm-1 and this was

assigned to 1-butene, which also was observed over V-P-O catalyst.71 Alternatively, the

peak at 1602 cm-1 can be assigned to the C=C mode of crotyl alcohol, while the peak at

1620 cm-1 could originate from the adsorbed 1,3-BD, as shown by Cavani et al.10,19 In one

of their experiments, however, this peak appeared when the surface temperature was

increased under inert flow, which, based on our experiments, will not give any higher C4

compounds in the vapor-phase. The peak at 1675 cm-1 is low enough in intensity that it

would be overwhelmed if adsorbed aldehydes were present on the surface. However,

Figure 4.6 shows that there was no peak was found at that wavenumber. This is possibly

Page 150: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

130

due to the depleted ethanol and/or the little bit of crotonaldehyde that was being made on

the surface. This argument is supported by the absence of any peaks at ~1700 cm-1,

indicating that there was not enough acetaldehyde on the surface to induce desorption and

aldol condensation. Our observation shows that the formation of acetaldehyde, being very

reactive to the WK (1:1) surface, leads to the formation of various C=C containing

intermediates, including surface enolate, polymerized acetaldehyde, and aldol

condensation products, i.e. crotonaldehyde and 2,4-hexadienal. The crotonaldehyde, does

not, however, proceed further to produce crotyl alcohol and 1,3-BD. The unsaturated

aldehyde tends to remain on the surface of the catalyst, which is why it is rarely observed

in the gas-phase.

2.6. DFT calculations ethanol, acetaldehyde, crotonaldehyde and crotyl alcohol

vibrational frequencies.

Disagreements in peak assignments arise mainly when interpreting the

acetaldehyde adsorption spectra, where polymerization on the catalyst surface takes place

at relatively low temperature,61,62 in addition to the desired aldol condensation8,59,65,72 and

Cannizzaro reaction.66,68 In ethanol adsorption spectra, an important peak at 1604 cm-1,

accompanied by a shoulder at 1581 cm-1, also sparks discrepancy in assignments.8,52,68 Peak

shifts are often observed, mainly in the C-O region, where the anchoring sites vary for

different catalysts and hence almost directly affect the vibration.13,14,30,53 We used DFT

structure optimized and calculated frequencies of both gas and adsorbed ethanol,

acetaldehyde, crotyl alcohol and crotonaldehyde molecules to aid in in-situ DRIFT spectra

assignments. In the calculation, a defect site on MgO (100), i.e. Mg3C2+O4C

2- (kink), was

used for adsorbate structural optimization. Another potential active site, the OH group on

Page 151: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

131

MgO, was considered since indirect correlation using XPS has pointed out that more OH

groups on the MgO leads to a better conversion.12 Our detailed analysis using in-situ

DRIFTS, however, showed that the OH groups’ involvement during the reaction is minimal

with its participations limited to substitutional chemisorption and reversible thermal

rearrangement. It is widely known that MgO-SiO2 has defect sites, kinks in particular, that

are stable up to the reaction condition,15,30,31,44 not to mention the recently synthesized MgO

catalyst that gives conversion as good as the SiO2-based material, which questions the

silica’s role on the reaction.12 The use of defect sites has been utilized before10,11,17 and

kink site was chosen as the active site based on the stability comparison that was carried

out by Chieregato et al.10 and Zhang et al.11 It is well known that vibrational frequencies

calculated using harmonic approximation are typically larger than the fundamental ones

observed experimentally with the various scaling factors typically used.73 We calculated

these scaling factors by optimizing ethanol, acetaldehyde, crotonaldehyde, and crotyl

alcohol molecules and comparing the frequencies with the experimental values.70 Using

the method of least-squares we determined the scaling factors to be 0.997, 0.9962, 0.9996,

and 0.9903 for ethanol, acetaldehyde, crotonaldehyde, and crotyl alcohol, respectively. The

scaling factors for all molecules are very close to unity and unscaled frequencies were used

to assign FTIR peaks. The complete scaling calculation can be found in the Supplementary

Information Table S4.1.

We performed frequency calculations for several permutations of ethanol,

acetaldehyde (enolate), crotonaldehyde, and crotyl alcohol adsorbed on a defected MgO

surface.17 The optimized structures are shown in Figure 4.10 and the calculated

frequencies are tabulated in Table 4.4. The two ethanol (I) species lead to acetaldehyde on

Page 152: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

132

Mg3CO4C and ethylene on Mg3CO5C, while there are three adsorption configurations of

acetaldehyde (II): physisorbed on Mg3C, chemisorbed on Mg3CO4C and enolate adsorbed

on Mg3CO4C. For crotonaldehyde, configuration (IV) due to the assumed facile dehydration

of acetaldol to crotonaldehyde was used.17 The crotyl alcohol (V) DFT calculation leads to

both the dissociated state and the coordinated 1,3-BD (VI) with the α-C still bound to the

O atom. Structure numbers, as shown in Figure 4.1, refer to the particular steps in the

ethanol catalytic transformation cycle. Acetaldol adsorption was not optimized in this

work due to its rare observation during experiments.1

DFT calculations show that ethanol can be adsorbed in two separate ways: via

dissociative adsorption and semi-dissociative adsorption, with the deprotonated ethoxy still

interacting strongly with the resulting surface hydroxyl. Typical adsorption spectra of

ethanol on an MgO surface show a peak at ~2700 cm-1 which was never previously

discussed.10,14,19 With the periodic DFT calculation, this peak can be assigned to the

stretching mode of the surface hydroxyl group. In Figure 4.6, vapor-phase acetaldehyde

did not appear in infrared spectra. Instead, a peak at 1653 cm-1 appeared at the intermediate

temperatures, and this can be correlated with the DFT-calculated C-O vibration at 1657

cm-1. This is accompanied by the build-up of C=C containing surface species, shown by

the peak at ~1600 cm-1 in Figure 4.6.

Acetaldehyde adsorption on the basic catalyst surface results in the infrared peaks

similar to those of crotonaldehyde (Figure 4.7 and 6). The experimental peak at 1382 cm-

1 was previously assigned to the -CH3 wagging mode of acetaldehyde and this may overlap

with the surface crotonaldehyde, owing to the similarities between Figure 4.7 and 6.59 The

complexity of this peak is demonstrated by the shoulders around it and its appearance in

Page 153: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

133

both Figure 4.7 and 6. To answer this question, we looked at crotonaldehyde frequency

calculation as well as the two acetaldehyde species and surface enolate (Figure 4.10). DFT

shows that four of the species all exhibited vibration around that wavenumber, which can

be seen in Table 4.4. Peak at 1284 cm-1 was also previously assigned to the C-O vibration

from acetaldehyde, and this also most likely overlaps with a vibrational mode from

physisorbed acetaldehyde.59 The peak at 1284 cm-1 is revealed to originate from -CHO

bending of an interacting surface species, in which the C=O bond is opened with the

molecule bridging two Mg atoms and one O atom, as shown in Figure 4.10-II. This

vibrational mode is not present on the physisorbed acetaldehyde, where the analog of that

vibration is presented as a peak at 1382 cm-1, contradicting assignment by Singh et al.59

Analysis of the C=C region inevitably presents the possibility that surface enolate is the

more reactive state of acetaldehyde. This enolate is the main building block for further

reactions, such as polymerization and aldol condensation.74 The presence of the enolate is

rather hard to confirm experimentally,74 due to its subsequent spontaneous reactions, but

DFT calculation reveals its presence, as verified the by C=C stretches at 1621 cm-1 and the

-CHO bending mode at 1384 cm-1. Finally, the peak at 2845 cm-1 in Figure 4.8 can be

assigned to the surface crotonaldehyde, blue-shifted from the vapor-phase. Our DFT

calculations show that this peak can be assigned to ν (CH), calculated at 2868 cm-1.

Page 154: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

134

I (Mg3CO4C) I (Mg3CO5C) Surface model

II (Mg3C) II (Mg3CO4C) II (Mg3CMg4C)

IV (Mg3CMg4C) V (Mg3CMg4C) VI (Mg3CMg4C)

Mg O

O4C Mg3C υCO = 1133 cm-1

υCH = 2914 cm-1

υCO = 1658 cm-1

υOH = 2678 cm-1

υCO = 1052 cm-1

υCC = 1098 cm-1

υCO = 1068 cm-1

υCC = 1621 cm-1

υCO = 1173 cm-1

υCC = 1661 cm-1

υCO = 1584 cm-1

υCC = 1663 cm-1

υCO = 1081 cm-1

υCC = 1587 cm-1

Figure 4.10. PBE optimized structures of ethanol (I), acetaldehyde (II), its enolate

conformation (II), crotonaldehyde (IV), crotyl alcohol (V) and 1,3-butadiene (VI) on MgO

surface low coordination Mg3cO4c or Mg3cO5c surface sites. Numbers refer to the particular

steps in catalytic transformation cycle shown in Figure 4.1.

Page 155: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

135

Table 4.4 Calculated infrared frequencies of ethanol, acetaldehyde, crotonaldehyde and

crotyl alcohol molecules adsorbed on low coordination model MgO surface sites.

Frequencies were calculated using PBE density functional and no scaling to correct for

anharmonicity was applied.

Assignment Ethanol Acetaldehyde Enolate Crotonaldehyde Crotyl

alcohol

Configuration Mg3CO4C Mg3CO5C Mg3C Mg3CO4C Mg3CMg4C Mg3CMg4C Mg3CMg4C

ν (OH) 3574 2679 - - 3063 - -

ν (CH3)

3030,

3020,

2950

3036,

3024,

2952

3096,

3019,

2965

3061,

3037,

2961

- 3000, 2943

3062,

3007,

2959

ν (CH2) 2903,

2873

2938,

2909 - -

3173,

3073 -

2946,

2875

ν (CH) - - 2914 2828 3028 3102, 3068,

3041, 2868

3056,

3040

ν (C=O) - - 1657 1016 1173 1664 -

ν (C=C) - - - - 1621 1581, 1010 1663

δ (CH2) 1460 1469 - - 1384 - 1378

δ (CH3) 1438,

1435

1449,

1440 - - - 1431, 1428 1350

ρw (CH) 1353,

1337 - - 1367 - 1232 -

ρw (CH2) - 1359 - - - - 1434

ρw (CH3) - 1341

1408,

1402,

1320

1447,

1428,

1320

- 1349 1441,

1419

ρw (CHO) - - 1382 1292 1302 1365 -

ρt (CH2) 1260 1267 - - - - -

δ (OH) - 1116 - - 1004 - -

δ (CC) - 1072 1117 1066 - - -

ν (CO) 1133 1052 - - - - -

ρw (CCO) 1132 - - - - - 1081

ν (CC) 1064 - - 1098 - - -

Page 156: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

136

The adsorption of crotyl alcohol provides much clarification for the entire reaction

sequence. Silica possesses weak Lewis acid sites and might provide an additional

dehydration site for the reaction. However, when ethanol was run on a bare silica catalyst,

only a little conversion of ethanol was achieved at higher temperature.75 The reaction

yielded acetaldehyde, as well as ethylene and ether, which explains silica’s role as a solid

acid catalyst. MgO has the ability to dehydrate the crotyl alcohol to give 1,3-butadiene,

which justifies the use of MgO defect sites for this reaction.1,75 The interaction with silica,

as shown in the Section 3.3, shows that the strong Lewis acid sites are due to the interaction

of silica and MgO.1,15 The peak at 1600 cm-1, which immediately forms on the surface

during experiment, indicates the presence of a C=C stretch. The vapor-phase crotyl alcohol

exhibits a vibration at around ~1670 cm-1,70 which is blue-shifted ~10 cm-1 for its adsorbed

state, based on the DFT calculation and shown in Table 4. These peaks are also observed

on the spectra in Figure 4.9, with the shifted peak being a shoulder to the intense, sharp

1602 cm-1. This indicates the presence of another surface species, which was shown by

DFT to be a coordinated 1,3-BD (Figure 4.10, VI). The assignment of this 1600 cm-1 peak

is also corroborated by the absence of the sp2 carbon C-H stretch peak, since α-C atom is

still in transition from sp3 to sp2.

2.7. In-situ DRIFT spectra for the ethanol, acetaldehyde, crotonaldehyde and

crotyl alcohol reaction on a WK (1:1) catalyst surface: the effect of the vapor phase

presence.

Formation of the C4 intermediates and products on MgO catalysts requires the

presence of the vapor phase ethanol and does not proceed via adsorbed ethoxide

intermediate catalytic conversion alone.19 This is also supported by our experiments, as

Page 157: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

137

discussed in Section 3.5, where ethanol desorption did not lead to the peaks caused by

acetaldehyde on the surface. Hence, we performed in-situ temperature programmed

DRIFTS experiments, during which a continuous reactant flow was carried out over a

sample containing adsorbed intermediates. DRIFT spectra for ethanol adsorbed on WK

(1:1) in the presence of a continuous vapor flow are shown in Figure 4.11. A spectrum of

the catalyst surface with the adsorbate at 373 K was used as a reference. The C-H stretching

region has peaks at 2978, 2933, 2903, and 2877 cm-1. These peaks are all attributed to CH3

and CH2 stretches. These peaks are also accompanied by the triplet at 1456, 1391 and 1061

cm-1 previously assigned to δ(CH3), δ(OH) and υ(CO) of vapor-phase ethanol.70,76 Two

other peaks can also be observed at 1630 and 1322 cm-1 in the low temperature 373 to 473

K regime, while higher temperatures result in their disappearance. While the latter can be

assigned to the wagging mode of the adsorbed ethoxy species, the former was previously

assigned to the adsorbed crotyl alcohol.19 It is highly unlikely, however, that crotyl alcohol

was being made at such a low temperature, since no acetaldehyde was observed. Hence,

we assigned the peak at 1630 cm-1 to adsorbed water in accordance with our assignment in

Section 3.5.1. The higher temperature regime, 523 to 723 K, also consistently resulted in

the greater 1575 and 1440 cm-1 peaks, in addition to peak broadening at ~3061 cm-1 which

appears to be stable on the catalyst surface even at these higher temperatures. The peak

broadening is indicative of the presence of olefins, i.e. ethylene and 1,3-BD. A notable

increase, shown by the 1743 and 1687 cm-1 peaks at intermediate temperatures, can be

assigned to acetaldehyde, both vapor-phase and chemisorbed, respectively. The C=O

stretch peaks coincide with the emergence of a 3004 cm-1 peak, which is attributable to the

C-H stretch of a sp2 carbon. These two peaks are accompanied by a broad band at 1280

Page 158: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

138

cm-1, which is the same with the peak at 1284 cm-1 in Figure 4.7. We have assigned this

peak to ρw of CHO. Right after the appearance of this intermediate, peaks at 1579 and 1434

cm-1 become very apparent, suggesting that these vibrations are from the reaction products

of adsorbed acetaldehyde. These two peaks are also similar to those observed in a similar

region, as shown in Figure 4.7 and Figure 4.8 for acetaldehyde and crotonaldehyde surface

adsorption, respectively. As previously assigned, these two peaks originate from a C=C

stretch (1579 cm-1) and the bending modes of CH2 or CH3 (1434 cm-1).

Page 159: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

139

Figure 4.11. In-situ DRIFTS spectra acquired of ethanol on WK (1:1) catalyst. Ethanol

was adsorbed on the sample surface, flown continuously and temperature increased from

376 to 723 K while spectra being recorded. In-situ DRIFTS spectrum of the sample

surface with adsorbed ethanol present at 373 K was used for reference.

Page 160: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

140

Complementary vapor-phase composition measurements performed using gas

chromatography (not shown) demonstrated that at a temperature regime above 523 K, 1,3-

BD was made but no crotonaldehyde and crotyl alcohol were observed. This suggests that

Figure 4.12. In-situ DRIFTS spectra acquired of acetaldehyde on WK (1:1) catalyst.

Acetaldehyde was adsorbed on the sample surface, flown continuously and temperature

increased from 376 to 723 K while spectra being recorded. In-situ DRIFTS spectrum of

the sample surface with adsorbed acetaldehyde present at 373 K was used for reference.

Page 161: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

141

various C4 intermediates to 1,3-BD tend to either react quickly with the product or remain

strongly adsorbed on the surface and react further, instead of desorbing. The correlation

made between the increase in 1,3-BD production by GC and the intensity increase in 1575

and 1440 cm-1 peak intensities implies that these two peaks are predominantly due to a C4

intermediate, with some contributions from surface enolate and polymerized acetaldehyde,

as also previously shown in Section 3.5.1. While the first peak was previously assigned to

2,4-hexadienal,8 surface acetates from the Cannizzaro reaction would exhibit both bands

due to their asymmetric and symmetric -COO stretching modes.52 Their increasing

intensity also suggests that the surface ethoxide, which dehydrogenates to acetaldehyde, is

continuously replenished by the vapor-phase ethanol, as shown in Figure 4.11. This

suggests that ethanol undergoes a catalytic transformation into acetaldehyde on WK (1:1),

which can be regarded as a rate limiting step because the other important intermediates are

spontaneously formed once acetaldehyde is produced. The subsequent aldol condensation

proceeds rapidly even at relatively low temperatures, as shown in Figures 4.5 and 4.6,

while crotyl alcohol is readily dehydrated, as demonstrated by the formation of 1,3-BD at

373 K.17,71 This observation agrees with literature reports, where on a basic catalyst with

little redox properties ethanol dehydrogenation is regarded as the rate limiting step.23,77

In-situ DRIFT spectra of acetaldehyde adsorbed on WK (1:1) in the presence of

vapor are shown in Figure 4.12. Peaks at 3060, 3023, 3000, 2962, 2931, 2870, 2821, 2791,

2736, and 2700 cm-1 are readily observed on these spectra. These various CH3, CH2, CH

sp2 stretches indicate the presence of multiple surface species. As discussed in Sections

3.5.1 and 3.5.2 this is due to the spontaneous polymerization, aldol condensation, and keto-

enol tautomerization of the acetaldehyde. A gradual increase in C-H stretching vibration

Page 162: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

142

from 2700 to 2821 cm-1 at lower temperatures is followed by their transformation into the

species responsible for the peaks at 2870 to 2962 cm-1 and can be associated with

transformation of vapor-phase acetaldehyde into chemisorbed acetaldehyde and into

surface enolate and crotonaldehyde. This is different from the temperature programmed

desorption experiments shown in Figure 4.7 and suggests that vapor-phase acetaldehyde

needs to be continuously supplied to replenish the surface species in order to continuously

produce C4 molecules. This experimental observation is also supported by the recent

computational study where aldol condensation of acetaldehyde on an MgO Mg3c site was

shown to result from the interaction between surface enolate and physisorbed

acetaldehyde.17 Peak assignments provided in Table 4.4 also suggest that peaks at 1762,

1724, 1442, 1343, and 1113 cm-1 can be assigned to the presence of vapor-phase

acetaldehyde. Higher temperature regimes above 573 K result in the decrease to 1724 cm-

1 as well as an enhanced 1273 cm-1 peak signifying the conversion of adsorbed

acetaldehyde into acetaldol as an intermediate that intermittently appears before being

dissociated to crotonaldehyde and water. The peak at 1273 cm-1 was previously observed

by Singh et al. and assigned to δ (C-OH) of the aldol.59 The peak at 1616 cm-1 is indicative

of the C=C stretch that originates from enolate, crotonaldehyde or 2,4-hexadienal. The peak

at 1650 cm-1 gradually increased with temperature, indicating the presence of adsorbed

aldehyde which could belong to acetaldehyde or crotonaldehyde. The triplet peak at ~1750

cm-1 slowly transforms into a singlet at higher temperatures indicating the depleted vapor-

phase acetaldehyde, due to the aldol condensation, confirmed by the presence of the

crotonaldehyde peak at 1700 cm-1. Similar to the case of ethanol, peaks at ~1574 and 1442

cm-1 increased with temperature. These peaks are, however, accompanied by a more

Page 163: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

143

prominent peak at 1555 cm-1. This peak has not been observed previously in Figures 4.5,

4.6 and 4.9. This ~20 cm-1 red shift is most likely due to the interaction between the C=C

molecule with vapor-phase acetaldehyde. In both ethanol and acetaldehyde reactive

desorption experiments in Figures 4.9 and 4.10 the same native hydroxyl groups at 3747,

3725, and 3680 cm-1 from Table 4.2 and Figure 4.3 are transiently involved in the catalytic

transformations. Furthermore, the hydroxyl group region appears to be similar to those in

Figure 4.5 suggesting no new basic sites are formed or they are immediately consumed by

the ensuing reactions. Figure 4.12 also suggests that aldol condensation on a WK (1:1)

surface proceeds quickly and is not a rate limiting step.

Figure 4.13. In-situ DRIFTS spectra acquired of crotonaldehyde on WK (1:1) catalyst

under ethanol vapor flow. Crotonaldehyde was adsorbed on the sample surface, flushed

with inert gas and ethanol was introduced under continuous flow with temperature

increased from 376 to 723 K while spectra being recorded. In-situ DRIFTS spectrum

of the sample surface with adsorbed crotonaldehyde at 373 K was used for reference.

For comparison, 523 K spectrum of crotonaldehyde adsorbed with no gas phase present

is shown in red dotted line.

Page 164: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

144

Meenvein-Ponndorf-Verley (MPV) reduction is a hydrogenation process in which

alcohols are used as a source of hydrogen.78 It was postulated to take place in the reaction

mechanism with ethanol hydrogenating the produced crotonaldehyde.1,16 It is typically

initiated by abstraction of an H+ from the alcohol. It was suggested that the rate limiting

step is hydride transfer from the adsorbed alcohol to the adsorbed carbonyl compound.78

Figure 4.13 shows the corresponding infrared spectra of the adsorbed crotonaldehyde in

the presence of ethanol vapor. Peaks at 2984, 2955, and 2902 can be specifically assigned

to ethanol CH3 and CH2 stretches with some minor contribution by other C4 molecules.

Negative peaks at 2845 and 2743 cm-1, on the other hand, are due to the CH3 Fermi

resonance and ν (CH) of crotonaldehyde, respectively.54,59 Each of these two negative

peaks is accompanied by a positive shoulder at a lower wavenumber, which also increases

with temperature, indicating the presence of another aldehyde, most likely acetaldehyde.

The presence of the vapor-phase acetaldehyde as the side product of the MPV reduction

can be confirmed by the C=O stretch at 1767 cm-1 , which has also been observed in

previous experiments under constant acetaldehyde and ethanol vapor flow. This

confirmation is shown in Figures 4.9 and 4.10. The peak at 1825 cm-1 can’t be assigned

to any of the alcohols or aldehydes. That peak can be found in the gas-phase 1,3-BD IR

spectrum, as reported in the NIST database.70 The signature C=C stretch of 1,3-BD is,

however, impossible to observe due to its overlap with the negative peaks from

crotonaldehyde. Negative peaks can readily be observed in the 1700-1600 cm-1 region and

1520-1400 cm-1 region, indicating the consumption of crotonaldehyde. A red-dotted IR

spectrum is shown as a comparison for the crotonaldehyde desorption at 523 K.

Comparison of the two spectra at the same temperature shows that the intensity decrease

Page 165: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

145

of the red plot is much less significant than when ethanol is constantly flown during the

experiment. The depletion of the surface crotonaldehyde is due to temperature-induced

desorption and reduction by ethanol to a certain extent. This intensity decrease is, however,

not accompanied by the peak at 1587 cm-1. This peak is relatively unaffected by the

temperature-programmed reaction, even though there are sharp peaks at 1560 cm-1, which

can also be found in the red-dotted spectrum. The absence of both positive and negative

peaks at 1587 cm-1 indicates that this peak is not part of the reactive intermediate and due

mostly to the overreaction of crotonaldehyde and acetaldehyde to 2,4-hexadienal.8 The

peak at 1520 cm-1 signifies the presence of a C-C containing molecule, which is being

consumed. Interestingly, peaks around 1460 and 1430 cm-1 are both initially consumed

before they start increasing positively. We expect this due to the possible overlap between

several CH3 containing molecules, such as crotonaldehye, which is initially consumed. The

acetaldehyde that is being produced and the vapor-phase ethanol, which is initially

consumed, starts to increase in intensity due to the depleted crotonaldehyde. The intensity

decrease in these peaks is also accompanied by the increasing intensity of the ethanol bands

at 1061, 1286, and 1346 cm-1.

3. Conclusions

Surface chemistry of WK (1:1) catalyst during the reaction of ethanol and the

corresponding reactive intermediates, including acetaldehyde, crotonaldehyde, crotyl

alcohol, has been investigated using in situ DRIFTS measurements combined with DFT

calculations. The nature of the native hydroxyl groups and their reactivity was also

investigated. They were found to undergo a transient reactivity via hydrogen bonded

interactions with the reactive intermediates. Ethanol adsorption resulted in several

Page 166: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

146

physisorbed and chemisorbed surface species. Acetaldehyde exhibited high reactivity to

yield crotonaldehyde but the excess resulted in strongly bound surface species assigned to

surface acetate, and/or 2,4-hexadienal or polymerized acetaldehyde. Crotonaldehyde is

more likely to be reduced by ethanol to yield crotyl alcohol than desorbing, even at

relatively high temperatures. Crotyl alcohol, on the other hand, showed to be very reactive

and adsorbs as two different species: physisorbed and deprotonated species that would

further desorb as 1,3-BD. Presence of gas phase hydrogen containing molecules, such as

ethanol, proved to be key in several reactive steps, including acetaldehyde condensation

step and crotonaldehyde reduction. Altogether, the data presented unraveled a complex

interplay between the surface hydroxyl groups, gaseous reactants and surface bound

reactive intermediates of 1,3-BD formation. These complex surface processes are depicted

in Figure 4.14. This elucidated surface reaction mechanism, combined with vapor-phase

intermediate characterization, can be used as a foundation for structure-activity relationship

study in combination with active sites determination. This will further lead to rational

design of catalyst. Future work will attempt to correlate vapor phase product evolution with

the most stable or transient reactive surface intermediates to examine trends leading to

higher overall 1,3-BD selectivity.

Page 167: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

147

(I)

(II)

(III) (IV)

(V)

(VI)

Figure 4.14. Complete surface reaction scheme on ethanol reaction over MgO/SiO2

catalyst. (I) Crotonaldehyde, (II) adsorbed crotyl alcohol, (III) 1,3-butadiene, (IV)

2,4-hexadienal, (V) paraldehyde, (VI) metaldehyde.

Page 168: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

148

Chapter 4 – Supporting Information

Figure S4.1. In-situ spectroscopy of ethanol on MgO catalyst. Ethanol was adsorbed

on the sample surface and temperature ramped up from 373 to 723 K while spectra

being recorded. Subtracted spectra are shown. Spectra are offset for clarity.

Figure S4.2. In-situ spectroscopy of acetaldehyde on MgO catalyst. Acetaldehyde was

adsorbed on the sample surface and temperature ramped up from 373 to 723 K while

spectra are being recorded. Spectra are offset for clarity.

Page 169: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

149

Table S4.1. Calculated infrared frequencies of gas phase ethanol, acetaldehyde, crotyl alcohol and crotonaldehyde molecules.

Frequencies were calculated using PBE density functional and no scaling to correct for anharmonicity was applied. Experimental

frequencies, except for crotonaldehyde, were obtained from NIST.68

Vibration

Ethanol Acetaldehyde Crotyl alcohol Crotonaldehyde

DFT Frequency

(cm-1)

IR

(cm-1)

DFT Frequency

(cm-1)

IR

(cm-1)

DFT Frequency

(cm-1)

IR

(cm-1)

DFT Frequency

(cm-1) IR (cm-1)

ν (OH) 3718 3686 - - 3702 3665 - -

ν CH3 3053, 3038,

2966

3035,

3012,

2960

3092, 3025, 2967

3024,

2996,

2967

3008, 2959 2970 3068, 3011,

2965, 2984

ν CH2 3011, 2925 3008,

2905 - - 3000, 2928

2940,

2880 - -

ν CH - - 2790 2840 3062, 3055, 3037 3030 3095, 3065,

3042, 2788

3044,

3007,

2750,

2828

ν (C=O) - - 1749 1743 - - 1691 1693

ν (C=C) - - - - 1675 1675 1644 1645

δ (CH2,

CH3) 1463 1456 1413, 1406

1410,

1390

1448, 1439,

1425, 1359

1480,

1435,

1415,

1338

1430, 1420,

1353

1449,

1398,

1381

δ (OH) 1325 1391 - - - - - -

δ (COH) - - - - - - 1367 Not

observed

149

Page 170: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

150

δ (CCH) 1237 1242 - - - - 1288 Not

observed

Combinati

on

bending

- - - -

1362, 1313,

1285, 1261,

1170, 1115, 1024

1384,

1290,

1250,

1180

1236 Not

observed

Combinati

on stretch - - - - 1075 1080 1142, 1086

1151,

1082

ν (CO) 1027 - - - 981 970 - -

ν C-C

(C=C) 863 - 1095 1122 - - - -

Scaling

factor 0.997 0.9962 0.9903 0.9996

150

Page 171: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

151

References

(1) Pomalaza, G.; Capron, M.; Ordomsky, V.; Dumeignil, F. Catalysts 2016, 6 (12),

203.

(2) Shylesh, S.; Gokhale, A. A.; Scown, C. D.; Kim, D.; Ho, C. R.; Bell, A. T.

ChemSusChem 2016, 9 (12), 1462–1472.

(3) Ostromislenskiy, J. J. Russ. Phys. Chem. Soc. 1915, No. 47, 1472 –1506.

(4) Lebedev, S. V. Zh Obs. Khim 1933, No. 3, 698–717.

(5) Posada, J. A.; Patel, A. D.; Roes, A.; Blok, K.; Faaij, A. P. C.; Patel, M. K.

Bioresour. Technol. 2013, 135, 490–499.

(6) Angelici, C.; Weckhuysen, B. M.; Bruijnincx, P. C. a. ChemSusChem 2013, 6 (9),

1595–1614.

(7) Janssens, W.; Makshina, E. V.; Vanelderen, P.; De Clippel, F.; Houthoofd, K.;

Kerkhofs, S.; Martens, J. A.; Jacobs, P. A.; Sels, B. F. ChemSusChem 2014, 8 (6),

994–1008.

(8) Ordomsky, V. V.; Sushkevich, V. L.; Ivanova, I. I. J. Mol. Catal. A Chem. 2010,

333 (1), 85–93.

(9) Muller, P.; Burt, S. P.; Love, A. M.; McDermott, W. P.; Wolf, P.; Hermans, I. ACS

Catal. 2016, 6 (10), 6823–6832.

(10) Chieregato, A.; Velasquez Ochoa, J.; Bandinelli, C.; Fornasari, G.; Cavani, F.;

Mella, M. ChemSusChem 2014, 8 (2), 377–388.

(11) Zhang, M.; Gao, M.; Chen, J.; Yu, Y. RSC Adv. 2015, 5 (33), 25959–25966.

(12) Hayashi, Y.; Akiyama, S.; Miyaji, A.; Sekiguchi, Y.; Sakamoto, Y.; Shiga, A.;

Koyama, T.; Motokura, K.; Baba, T. Phys. Chem. Chem. Phys. 2016, 18 (36),

25191–25209.

(13) Hanspal, S.; Young, Z. D.; Shou, H.; Davis, R. J. ACS Catal. 2015, 1737–1746.

(14) Birky, T. W.; Kozlowski, J. T.; Davis, R. J. J. Catal. 2013, 298 (0), 130–137.

(15) Kvisle, S.; Aguero, A.; Sneeden, R. P. A. Appl. Catal. 1988, 43 (1), 117–131.

(16) Toussaint, W. J.; Dunn, J. T.; Jachson, D. R. Ind. Eng. Chem. 1947, 39 (2), 120–

125.

(17) Taifan, W. E.; Bučko, T.; Baltrusaitis, J. J. Catal. 2017, 346, 78–91.

(18) Quattlebaum, W. M.; Toussaint, W. J.; Dunn, J. T. J. Am. Chem. Soc. 1947, 69,

593–599.

(19) Velasquez Ochoa, J.; Bandinelli, C.; Vozniuk, O.; Chieregato, A.; Malmusi, A.;

Recchi, C.; Cavani, F. Green Chem. 2016, 18, 1653–1663.

(20) Velasquez Ochoa, J.; Malmusi, A.; Recchi, C.; Cavani, F.; Ochoa, J. V.; Malmusi,

A.; Recchi, C.; Cavani, F. ChemCatChem 2017, 9 (12), 2128–2135.

(21) Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A.

ChemSusChem 2014, 7 (9), 2505–2515.

(22) Makshina, E. V.; Janssens, W.; Sels, B. F.; Jacobs, P. A. Catal. Today 2012, 198

(1), 338–344.

(23) Makshina, E. V; Dusselier, M.; Janssens, W.; Degreve, J.; Jacobs, P. A.; Sels, B. F.

Chem. Soc. Rev. 2014, 43 (22), 7917–7953.

(24) Wu, C.-M.; Baltrusaitis, J.; Gillan, E. G.; Grassian, V. H. J. Phys. Chem. C 2011,

115 (20), 10164–10172.

(25) Haas, P.; Tran, F.; Blaha, P. Phys. Rev. B 2009, 79 (8), 85104.

Page 172: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

152

(26) Kantorovich, L. N.; Holender, J. M.; Gillan, M. J. Surf. Sci. 1995, 343 (3), 221–

239.

(27) Chizallet, C.; Petitjean, H.; Costentin, G.; Lauron-Pernot, H.; Maquet, J.;

Bonhomme, C.; Che, M. J. Catal. 2009, 268 (1), 175–179.

(28) Chizallet, C.; Bailly, M. L.; Costentin, G.; Lauron-Pernot, H.; Krafft, J. M.; Bazin,

P.; Saussey, J.; Che, M. Catal. Today 2006, 116 (2), 196–205.

(29) Di Valentin, C.; Del Vitto, A.; Pacchioni, G.; Abbet, S.; Wörz, A. S.; Judai, K.;

Heiz, U. J. Phys. Chem. B 2002, 106 (46), 11961–11969.

(30) Branda, M. M.; Rodríguez, A. H.; Belelli, P. G.; Castellani, N. J. Surf. Sci. 2009,

603 (8), 1093–1098.

(31) Bailly, M.; Chizallet, C.; Costentin, G.; Krafft, J.; Lauronpernot, H.; Che, M. J.

Catal. 2005, 235 (2), 413–422.

(32) Baltrusaitis, J.; Hatch, C.; Orlando, R. J. Phys. Chem. A 2012, 116 (30), 7950–

7958.

(33) Kresse, G.; Hafner, J. Phys. Rev. B 1994, 49 (20), 14251–14269.

(34) Kresse, G.; Furthmüller, J. Comput. Mater. Sci. 1996, 6 (1), 15–50.

(35) Kresse, G.; Furthmüller, J. Phys. Rev. B 1996, 54 (16), 11169–11186.

(36) Kresse, G.; Hafner, J. Phys. Rev. B 1993, 48 (17), 13115–13118.

(37) Blöchl, P. E. Phys. Rev. B 1994, 50 (24), 17953–17979.

(38) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77 (18), 3865–3868.

(39) Knözinger, E.; Jacob, K.-H.; Singh, S.; Hofmann, P. Surf. Sci. 1993, 290 (3), 388–

402.

(40) Coluccia, S.; Lavagnino, S.; Marchese, L. Mater. Chem. Phys. 1988, 18 (5), 445–

464.

(41) Anderson, P. J.; Horlock, R. F.; Oliver, J. F. Trans. Faraday Soc. 1965, 61 (0),

2754–2762.

(42) Shido, T.; Asakura, K.; Iwasawa, Y. J. Chem. Soc.{,} Faraday Trans. 1 1989, 85

(2), 441–453.

(43) Morrow, B. A. Stud. Surf. Sci. Catal. 1990, 57, A161–A224.

(44) Chizallet, C.; Costentin, G.; Che, M.; Delbecq, F.; Sautet, P.; Surface, D.; Marie,

P. J. Am. Chem. Soc. 2007, No. 19, 6442–6452.

(45) Hair, M. L. J. Non. Cryst. Solids 1975, 19, 299–309.

(46) Chung, S.-H.; Angelici, C.; Hinterding, S. O. M.; Weingarth, M.; Baldus, M.;

Houben, K.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ACS Catal. 2016, 6 (6),

4034–4045.

(47) Hofmeister, A. M.; Bowey, J. E. Mon. Not. R. Astron. Soc. 2006, 367 (2), 577–

591.

(48) Mellini, M. Am. Mineral. 1982, 67 (5–6), 587–598.

(49) Taifan, W.; Boily, J.-F.; Baltrusaitis, J. Surf. Sci. Rep. 2016, 71 (4), 595–671.

(50) Vimont, A.; Thibault-Starzyk, F.; Daturi, M. Chem. Soc. Rev. 2010, 39 (12), 4928–

4950.

(51) Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A. Catal.

Sci. Technol. 2015.

(52) Song, H.; Ozkan, U. S. J. Catal. 2009, 261 (1), 66–74.

(53) Phung, T. K.; Lagazzo, A.; Rivero Crespo, M. Á.; Sánchez Escribano, V.; Busca,

G. J. Catal. 2014, 311, 102–113.

Page 173: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

153

(54) Raskó, J.; Kiss, J. Appl. Catal. A Gen. 2005, 287 (2), 252–260.

(55) Busca, G.; Lorenzelli, V. Mater. Chem. 1982, 7 (1), 89–126.

(56) Wang, H.; Schneider, W. F. Phys. Chem. Chem. Phys. 2010, 12 (24), 6367–6374.

(57) Meixner, D. L.; Arthur, D. A.; George, S. M. Surf. Sci. 1992, 261 (1), 141–154.

(58) Raskó, J.; Dömök, M.; Baán, K.; Erdőhelyi, A. Appl. Catal. A Gen. 2006, 299,

202–211.

(59) Singh, M.; Zhou, N.; Paul, D. K.; Klabunde, K. J. J. Catal. 2008, 260 (2), 371–

379.

(60) Cederstav, A. K.; Novak, B. M. J. Am. Chem. Soc. 1994, 116 (9), 4073–4074.

(61) Georgieff, K. K. J. Appl. Polym. Sci. 1966, 10 (9), 1305–1313.

(62) Bevington, J. C.; Norrish, R. G. W. Proc. R. Soc. London. Ser. A. Math. Phys. Sci.

1949, 196 (1046), 363 LP-378.

(63) Cannizzaro, S. Justus Liebigs Ann. Chem. 1853, 88 (1), 129–130.

(64) Tishchenko, V. E. J. Russ. Physico-Chemical Soc. 1906, 38 (355–418).

(65) Diaz, C. D. C.; Locatelli, S.; Gonzo, E. E. Zeolites 1992, 12, 851–857.

(66) Idriss, H.; Diagne, C.; Hindermann, J. P.; Kiennemann, A.; Barteau, M. A. J.

Catal. 1995, 155 (2), 219–237.

(67) Guil, J. M.; Homs, N.; Llorca, J.; De La Piscina, P. R. J. Phys. Chem. B 2005, 109

(21), 10813–10819.

(68) Resini, C.; Cavallaro, S.; Frusteri, F.; Freni, S.; Busca, G. React. Kinet. Catal. Lett.

2007, 90 (1), 117–126.

(69) Ho, C. R.; Shylesh, S.; Bell, A. T. ACS Catal. 2016, 6 (2), 939–948.

(70) Stein, S. E. In NIST Chemistry WebBook, NIST Standard Reference Database

Number 69; Linstrom, P. J., Mallard, W. G., Eds.; National Institute of Standards

and Technology (NIST): Gaithersburg, MD, 2017.

(71) Wenig, R. W.; Schrader, G. L. J. Phys. Chem. 1987, 91 (22), 5674–5680.

(72) Rekoske, J. E.; Barteau, M. a. Langmuir 1999, 15 (6), 2061–2070.

(73) Scott, A. P.; Radom, L. J. Phys. Chem. 1996, 100 (41), 16502–16513.

(74) Palagin, D.; Sushkevich, V. L.; Ivanova, I. I. J. Phys. Chem. C 2016, 120 (41),

23566–23575.

(75) Sekiguchi, Y.; Akiyama, S.; Urakawa, W.; Koyama, T.; Miyaji, A.; Motokura, K.;

Baba, T. Catal. Commun. 2015, 68, 20–24.

(76) Plyler, E. K. J. Res. Natl. Bur. Stand. (1934). 1952, 48 (4).

(77) Niiyama, H.; Morii, S.; Echigoya, E. Bull. Chem. Soc. Jpn. 1972, 45 (3), 655–659.

(78) Hattori, H. Chem. Rev. 1995, 95 (3), 537–558.

(79) Towns, J.; Cockerill, T.; Dahan, M.; Foster, I.; Gaither, K.; Grimshaw, A.;

Hazlewood, V.; Lathrop, S.; Lifka, D.; Peterson, G. D.; Roskies, R.; Scott, J. R.;

Wilkins-Diehr, N. Comput. Sci. Eng. 2014, 16 (5).

Page 174: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

154

Chapter 5

Active Sites Determination of MgO/SiO2

Catalysts for Ethanol to 1,3-BD Reaction

Abstract .......................................................................................................................... 154

1. Introduction ....................................................................................................... 155

2. Results and Discussion ...................................................................................... 160

2.1. Steady state ethanol catalytic conversion to 1,3-BD ................................. 160

2.2. Bulk, surface chemical and structural characterization using XRD, LEIS

and DRIFTS ........................................................................................................ 163

2.3. Temperature-programmed reaction spectroscopy (TPRS) of ethanol on

MgSi-WK ............................................................................................................ 167

2.4. Acid-base characterization using DRIFTS ................................................ 172

2.5. Reactive site persistence during ethanol-to-1,3-BD .................................. 174

2.6. Implications for the structure-activity relationship .................................... 186

3. Conclusions ........................................................................................................ 189

References ...................................................................................................................... 191

Abstract

Ethanol is an important renewable chemical that allows for sustainable high value

product, such as 1,3-butadiene, catalytic synthesis. MgO/SiO2 catalyst is typically utilized

in a single step ethanol-to-1,3-butadiene catalytic conversion and the (by)product yields

were shown to depend on the type, structure and strength of the catalytic active sites. The

fundamental factors describing the molecular structure and binding properties of these sites

is thus of critical importance but not yet fully understood. We utilized multimodal

approach, including temperature programmed surface sensitive infrared, mass

Page 175: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

155

spectroscopy using probe molecules, such as CO2, NH3, pyridine and propionic acid, to

unravel the structure and persistence of these catalytic sites in situ. In particular, Mg-O-

Mg, Mg-O(H)-Mg, Mg-O-Si and Mg-O(H)-Si surface site binding configurations were

interrogated using spectroscopic methods in combination with DFT calculations. Surface

elemental analysis using low energy ions suggested that either Mg atoms or Si being the

most abundant on the topmost surface layer, depending on the catalyst preparation method.

The molecular active site structure was determined and incipient wetness prepared surface

was found to be dominated with stabilized Mg-OH with little magnesium silicate (Mg-O-

Si and Mg-O(H)-Si) functional groups. The wet-kneaded catalyst surface, on the other

hand, contained a significant number of surface sites derived from magnesium silicates.

The fundamental surface site structure proposed here can further serve as a starting point

for theoretical calculations necessary to fully model the reactive pathway during ethanol

catalytic transformation to 1,3-butadiene.

1. Introduction

Elucidating surface active site structure is of high importance for development of

selective MgO/SiO2 catalysts utilized for the catalytic conversion of ethanol to 1,3-

butadiene (1,3-BD). Deemed as Lebedev catalyst, it is increasingly investigated due to its

bifunctional nature and the variety of the (by)products that can be formed thereon.1–4 The

nature of the closed shell MgO electronic structure and diverse surface functionality

provides interesting challenges as there are much fewer spectroscopic methods that allow

identification of the active site properties, akin to those of solid Lewis acid catalysts, such

Page 176: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

156

as Ta2O55 or ZrO2,

6 also used in ethanol-to-1,3-butadiene catalytic conversion (1,3-BD).3,7

As a result, the surface site structure and the reactivity of MgO/SiO2 catalysts during 1,3-

BD production are still poorly understood.8 Furthermore, discrepancies between the

selectivity of the reported active catalysts are due to the intrinsic active site density, their

functional nature (acidic or basic) and their strength which arise from the diverse set of

preparation methods including ratio of Mg-to-Si, Mg precursor used and their deposition

method. The ratio of acidic-to-basic sites was shown to affect the overall reactivity during

1,3-BD formation, as demonstrated by Angelici et al.9 Shylesh et al. proposed that weak

basic sites were responsible for ethanol dehydrogenation and other basic sites for aldol

condensation.3 Angelici et al. attributed higher overall reactivity to a small number of

strong basic sites on the catalyst surface with an intermediate amount of acidic sites and

weak basic sites.9 Catalysts prepared using different methods, e.g. incipient wetness

impregnation (IWI) and wet-kneading (WK)3,9 resulted in a large activity difference with

IWI prepared catalyst yielding only ~5% conversion at 300 °C when compared to WK

catalyst ~50% conversion at 425 °C. The key to this different activity was believed to be

the significant improvement of acetaldehyde production by transition metal promotion on

strong basic sites.3 In general, MgO/SiO2 wet-kneading has consistently been shown to

produce highest 1,3-BD yields4,10,11 due to the proposed balanced acidic and basic catalytic

site number.9 The exact molecular structure of these acidic and basic active sites is still

under debate with most analysis focused correlating the reactivity and the (bulk) crystalline

catalyst phases.1,2,12 In particular, -Mg-O-Si- linkage has been implicated to be reactive and

related to the selectivity of the catalysts.1–3 SiO2 was proposed to indirectly catalyze the

reaction due to its structural perturbation of MgO using wet-kneading.10 Furthermore, a

Page 177: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

157

solid solution of MgO with SiO2 was inferred from the experimental measurements and

the amount of magnesium silicate phases, measured by 1H-29Si cross-polarized MAS

NMR1,3 and DRIFTS,2 was correlated to varying overall selectivity. Ochoa et al. observed

formation of a magnesium silicate phase with Mg2+ neighbored by Si4+ cations synthesized

using sol-gel methods with Mg/Si ratio of 9 to 15 while the lower ratio led to the formation

of catalytically inactive forsterite (Mg2SiO4) phase formation.2 Amorphous magnesium

silicate hydrous phase formed during wet-kneading of MgO and SiO2 was found to be

responsible for ethylene byproduct formation while layered magnesium silicate hydrous

phase was correlated to the 1,3-BD product.1 Furthermore, the silicate-to-MgO ratio was

suggested to be the key to the appropriate balance of acidic-basic sites.1 A different view

was offered by Shylesh et al. where hydroxyl (OH) groups were necessary in the proximity

of the strong basic Mg2+-O2- sites to synergistically catalyze the reaction. Finally, the

correlation between the magnesium silicate hydrous phase and 1,3-BD yield was

challenged by Hayashi et al. who reported MgO catalyst that did not require participation

from SiO2 for this reaction.12 Said MgO catalyst was synthesized with an additional

hydrothermal step using NH4OH solution. XPS characterization of the two different MgO

catalysts, i.e. with and without the additional hydrothermal step, showed that latter, i.e. the

more active catalyst, exhibited a higher intensity of an unassigned O1s oxygen peak at

around 532 eV.12 The presence of this unidentified oxygen species on MgO could be related

to the reactive lower-coordinated oxygen atoms on Mg-O defect sites.13–15 Concurrently,

these lower-coordinated Mg-O pairs (Mg2+3CO2-

4C, Mg2+3CO2-

3C, Mg2+4CO2-

4C) were

computationally shown to be involved in 1,3-BD formation from ethanol.16–18 Analysis of

Lewis acid - ZrO2-based catalysts19 - suggests that Lewis acid sites (LAS) can be chiefly

Page 178: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

158

responsible for the activity in this reaction. By definition, closed Lewis acid heteroatoms

(M) are tetrahedrally coordinated (M-(OSi)4 to the zeolite framework, while open Lewis

acid heteroatoms are tri-coordinated (HO)-M-(OSi)3 to the zeolite framework.20–22 With

this in mind, the octahedral symmetry in MgO crystal12,23 allows to identify several LAS

as part of the intrinsic acid/base pairs to be available. i.e. Mg-O-Mg, Mg-O(H)-Mg, Mg-

O-Si and Mg-O(H)-Si. These combinations can further exist in open and closed acid

configurations, where the oxygen is bound to SiO2 while also coordinated to a proton to

form coordinated hydroxyl groups.24,25 Strict terminology of the open acid site requires an

isolated hydroxyl group to be present and while it is very basic, this hydroxyl group

spectroscopically has been proven to be non-existent.25,26 In addition to these sites, the

coordination of Mg is also very important, since catalysis by this metal oxide is driven by

defect sites.15,27,28 These proposed catalytic sites are shown in Figure 5.1.

While Zr-based catalysts mainly concerns the Zr-coordination into the framework,

and consequently, characterization of the resulting LAS, i.e. open and closed,19,29 study on

MgO/SiO2-based catalysts mostly revolves around the general acidity and basicity

characterization while also discussing the importance of bulk silicate phases.1,9,30,31

However, pyridine-DRIFTS studies concluded LAS to be the only acidic sites on the

catalyst as demonstrated by the IR peaks at 1450, 1578, and 1612 cm-1.3,9,30,31 NH3-TPD,

another routinely used acidity probe method, also discriminates the acid based on the

strength, without discussing the nature of the acidic sites.9 Hence the molecular structure

of these acidic sites is not well known.1 In this work, we combined spectroscopic

measurements in-situ using different probe molecules to identify the role of each sites

during the reaction and elucidate their molecular coordination. In particular, we begin by

Page 179: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

159

performing bulk XRD, surface LEIS and DRIFTS analysis of native surface hydroxyl

groups. We then perform steady state and kinetic temperature programmed experiments

of ethanol conversion to 1,3-BD using catalysts synthesized with different methods. We

then utilize temperature programmed DRIFTS to explore surface acidic and basic site

structure with ab initio calculations to support our NH3 adsorption site assignments and

hence propose the molecular arrangements of the catalytic sites. Sodium (Na) poisoning

was utilized to elucidate the role of the acidic sites during the reaction, which will further

indicate the importance of strong acidic sites during the reaction. Finally, the persistence

of these reactive sites is probed spectroscopically under the relevant conditions of

temperature and ethanol vapor.

B

C

D

E F

A

Figure 5.1. Possible combination of metal atoms that act as Lewis acid sites: A: Mg3C

(open), B: Mg3C (closed), C: Mg4C (closed), D: Mg4C (open), E: Mg5C (open), F: Mg5C

(closed).

Page 180: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

160

2. Results and Discussion

2.1. Steady state ethanol catalytic conversion to 1,3-BD

Steady state reactivity of the synthesized MgO/SiO2 catalysts was investigated

using fixed-bed reactor. A catalyst calcined at 500°C was synthesized as a benchmark

catalyst. This benchmark calcination temperature was based on earlier study by Zhu, et al.,

where 500°C was the optimized calcination temperature that resulted in a balanced acidic-

basic sites at 40.8 µmol/g v 49 µmol/g.32 Activity comparison was performed at a

temperature of 450 °C with the maximum 1,3-BD yield. At this reaction temperature,

carbon balance for each catalyst was determined to be >80%. Table 5.1 shows that ethylene

selectivity was the highest for MgSi-WK2 which suggests the presence of acidic sites on

the surface since ethanol dehydration reaction is very prominent over catalysts with very

high density of BAS.33–35 For catalysts calcined at 800 °C (MgSi-WK and MgSi-IWI)

ethylene selectivity was above 50%, which is intriguing, since pyridine probing does not

show the presence of BAS (vide infra). This ethylene formation can be proposed due to the

reactive Mg-O-Mg or Mg-O-Si linkages that are inherent in the catalysts.36 In agreement,

DFT calculations have shown that ethanol dehydration competes with dehydrogenation

reaction over LAS in MgO catalysts.16

Comparison between two catalysts calcined at 800 oC, i.e. MgSi-WK and MgSi-

IWI, shows that MgSi-WK was more active and selective to 1,3-BD suggesting that the

preparation method deeply affected the balance of the active sites. The accumulation of

acetaldehyde on MgSi-IWI catalyst was evident from Table 1 suggesting the active sites

for further aldol condensation and MPV (Meerwein-Ponndorf-Verley) reduction –

mechanistic steps taking place after ethanol dehydrogenation - were limited.16 MgSi-WK,

Page 181: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

161

on the other hand, exhibited significantly higher 1,3-BD selectivity and limited

acetaldehyde production suggesting more sites available for the subsequent reactions.

Figure 5.2 shows the evolution of each major (by)product with increasing temperature for

reaction over MgSi-WK catalyst. 1,3-BD and ethylene exhibited almost linear increase in

productivity with the calculated apparent activation energy of 12.4 and 18.02 kcal/mol,

respectively. Importantly, Arrhenius plot shown in Figure 5.2 inset of acetaldehyde

formation did not show linearity due to its involvement into further reactions.

Table 5.1 Steady state reactivity of MgO/SiO2 catalysts of different calcination

temperature and preparation method. Reaction was carried out at 450 °C with catalyst

mass of 0.1 g, 55 ml/min total flow rate and pethanol = 2.5 kPa. Selectivity towards major

(by)products ethylene, acetaldehyde and 1,3-BD is reported.

Catalyst Selectivity (%) Conversion

(%) Ethylene Acetaldehyde 1,3-BD

MgSi-WK 55.8 14.4 29.7 77.0

MgSi-WK2 82.9 9.7 7.5 60.4

MgSi-IWI 58.0 25.6 16.4 63.9

Productivity values of MgSi-WK compared reasonably well with those found in the

literature. In the present work 1,3-BD yield translated to the production rate of 0.44 gBD.gcat-

1.hr-1, while the MgSi-WK2 yielded about 0.06 gBD.gcat

-1.hr-1. Chung et al. synthesized a

WK catalyst using calcination temperature of 500 °C with the reported 1,3-BD productivity

of 0.23 gBD.gcat-1.hr-1.1 The origin of this reactivity can be related to the high amount of

layered hydrous magnesium silicate phase on the catalyst, which was highly dependent on

the precursor; a nano-sized Mg(OH)2 precursor was the preferred precursor.1 The effect of

MgO precursor used during the WK catalyst synthesis has recently been highlighted by

Huang et al. where Mg(OH)2 precursor synthesized using a template method yielded a very

high productivity of 1.15 gBD.gcat-1.hr-1.4 Since precursor was not the controlled variable in

Page 182: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

162

the present work, the origin of higher activity in the case of MgSi-WK, as compared to

MgSi-WK2, is presumably from the higher calcination temperature. The effect of the

calcination temperature has been observed previously for Zr/SiO2 catalysts where highest

1,3-BD yield was obtained with catalyst calcined at 550 °C.37 Study on the effect of

calcination temperature on MgO/SiO2 catalysts, on the other hand, revealed linear increase

in basicity with increasing calcination temperature.38 Pyridine testing indicated that

calcination temperature of 500 °C resulted in a catalyst that exhibited the highest acidity.38

Incipient wetness impregnation (IWI) method is not typically used for the MgO/SiO2

catalysts synthesis for this reaction despite its popularity in supported catalyst synthesis.

Typically, IWI method is utilized to obtain sub-monolayer coverage to prevent the

formation of second layer of bulk oxides. The activity of this catalyst was very different

yielding a much lower 1,3-BD productivity. Shylesh et al. investigated a similar catalyst

calcined at 500 °C and reported rather low 1,3-BD yield of 0.01 gBD.gcat-1.hr-1 at 300 °C.3

Surface layer compositions of this catalyst as well as MgSi-WK will further be interrogated

using LEIS to better understand chemical composition changes leading to such different

reactivity.

Importantly, steady state experiments suggest that ethylene and acetaldehyde are

the most encountered byproducts during the reaction. As shown in Table 5.1, their

combined selectivity makes up to more than 70% of the total activity. While acetaldehyde

recycling can be utilized due to it being a reactive intermediate, ethylene production should

be limited. One of the advantages of using MgO/SiO2 is no butene byproduct production.

Butenes form azeotrope with 1,3-BD and increase any separation cost. It was reported that

butenes could be formed by hydrogenation of 1,3-BD over platinum catalysts, dehydration

Page 183: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

163

of n-butanol, and from thermal or catalytic dimerization of ethylene.39 Hence n-butanol

dehydration might not be feasible considering the high selectivity of ethylene. However,

although the reactive pathway is similar, n-butanol synthesis requires a tautomerization site

to convert crotyl alcohol to 1-buten-1-ol over a basic oxygen site, which further requires

another MPV reduction site.40 This pathway is not supported by our catalytic, DRIFTS36

and TPRS data (vide infra) since no n-butanol was formed in the product stream and no

butyraldehyde was spectroscopically observed.

2.2. Bulk, surface chemical and structural characterization using XRD, LEIS and

DRIFTS

Bulk crystalline structure of the catalysts was characterized using XRD. XRD

patterns of the selected MgSi-WK catalysts as a function of the corresponding oxide ratio

as well as those for MgSi-IWI are compared in Figure 5.3. XRD pattern of WK catalysts

indicates the formation of periclase MgO in the bulk. The intensity expectedly enhanced

Figure 5.2. Catalytic activity of MgSi-WK between 350-450°C. Inset: Arrhenius plot

of ethylene and 1,3-BD. Catalyst mass = 0.1 gr, total flow rate = 55 ml/min, pethanol =

2.5 kPa.

Page 184: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

164

with higher Mg content which means that the catalyst crystallinity originated from MgO

rather than any silicate material. On the other hand, IWI catalyst showed a very different

crystalline structure. Several crystalline phases were identified in the XRD patterns

including those that might be attributed to periclase phase of MgO. However, most of the

peaks were due to the presence of different crystalline phase, potentially magnesium

silicates.41–44 While this provided implications for the varied selectivity shown in Table

5.1, surface chemical analysis was performed to further elucidate this effect.

Topmost surface layer of the two catalysts, i.e. MgSi-IWI and MgSi-WK, was

probed using LEIS. LEIS is by far the most surface sensitive characterization technique

which sputters the surface with very low energy ions using ionized noble gases.45 LEIS

spectra of both catalysts are shown in Figure 5.4. Sputtering experiments (depth profiling)

are shown as insets where surface layers were sputtered using 1 keV Ar+ ions. The

sputtering rates were on the order of one monolayer of atoms per 1015 Ar+ ions/cm2. The

legend in the insets indicates the nth layer removed from the catalyst surface. Three peaks

Figure 5.3. Comparison of XRD patterns of MgSi-WK and MgSi-IWI. WK with

different oxide ratios are overlaid for comparison.

Page 185: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

165

were found in the spectra corresponding to oxygen at ~1200 eV, Mg at ~1650 eV and a

shoulder for Si at ~1760 eV. The increase in signal intensity for all peaks after initial

sputtering may be due to some residual overlayer on the surface or perhaps to initial

planarization of the catalyst granules increasing the apparent global atomic surface density.

Depth profile spectra were obtained stepwise using the acquisition of the surface spectra

followed by sputtering for a designated period. The resulting depth profiles show Mg-rich

surface for MgSi-IWI while Si-rich surface can be for MgSi-WK. The incipient wetness

impregnation method deposited the magnesium nitrate precursor on top of the fumed SiO2

yielding a high magnesium content. The wet-kneaded method, on the other hand, provided

intimate mixing between the Mg(OH)2 and SiO2 allowing for the extensive interaction

between the two oxides which is reflected in the abundance of Si on the surface.1 Based on

this characterization, MgSi-IWI should contain more Mg-O-Mg linkages, while MgSi-WK

would contain more Mg-O-Si linkages.

The structure of the native OH groups of all the catalysts was investigated using

DRIFTS (Figure 5.5). MgSi-WK shows four major peaks in the spectrum in addition to a

Figure 5.4. Depth-profile of (a) MgSi-IWI and (b) MgSi-WK as probed using HS-LEIS.

HS-LEIS spectra of layer by layer sputtering of catalyst surface are shown in the inset.

Page 186: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

166

broad peak at ~3550 cm-1 assigned to germinal and vicinal OH groups of the silica

support.46,47 Peak at 3745 cm-1 was assigned to both isolated silanol group of the SiO2

support46,47 which decreased in intensity when MgO was wet-kneaded and to an isolated

OH group of MgO that depends on the coordination number of the Mg.36 An intense peak

at 3680 cm-1 was previously assigned to a magnesium silicate phase, lizardite.1 The other

peaks at 3725 and 3705 cm-1 were assigned to the isolated O4c-H and O5c-H coordinated

groups formed in the presence of the amorphous SiO2 (SiMg4cO4c and SiMg4cO5c).36

Formation of the peaks at 3725 and 3705 cm-1 was also confirmed with varying Mg/Si ratio

as shown in Figure 5.5 (right). At very high Mg content, i.e. Mg/Si > 7/3, it can be seen

that a very basic Mg-OH group started to become apparent at 3765 cm-1. As previously

shown,36 MgO possessed two basic OH groups that depend on the coordination number of

oxygen atoms with the lower coordinated OH group at higher wavenumbers.25,48

Interestingly, the peaks at 3705 and 3680 cm-1 increased with Mg content while the peak

at 3725 cm-1 intensified at intermediate ratio and diminished at both extremes. This

suggests that the peaks at 3680 and 3705 cm-1 were dominant at lower Si content but not

in pure MgO, and hence, it can be assigned to an OH group anchored to MgSi coordination

that has few Si atoms nearby. The formation of this peak confirms the XRD assignment

where the catalyst is becoming increasingly crystalline MgO-like, both on the surface and

in the bulk.

The OH groups that were observed on MgSi-IWI surface were very different from

those found on MgSi-WK. Two new peaks at 3610 and 3571 cm-1, in addition to the peaks

that were also found on MgSi-WK, appeared, indicating the formation of two entirely new

OH groups. From HS-LEIS experiment, the top surface layer mostly consisted of Mg

Page 187: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

167

atoms, which would suggest the presence of a lot of Mg-OH groups. As previously

explained, the isolated OH groups from MgO exhibit peaks at both ~3760 and ~3740 cm-

1, depending on the coordination number. The presence of SiO2 likely stabilized the lowest-

coordinated OH group, i.e. peak at ~3760 cm-1, converting it into the higher-coordinated,

isolated OH group, i.e. peak at ~3740 cm-1. The other peaks at 3610 and 3571 cm-1 are

sharp with lower intensity, unlike the broad OH peaks that were assigned to multi-

coordinated hydrogen-bonded OH groups,24,25 These two peaks also were observed by

Ochoa et al. in the catalysts structurally similar to forsterite.2 From HS-LEIS, XRD, and

DRIFTS, the surface of MgSi-IWI was confirmed to be mainly populated by Mg-OH, with

significantly less magnesium silicate phases. Most of the formed magnesium silicate

phases were found in the bulk of the catalyst as indicated by the XRD.

2.3. Temperature-programmed reaction spectroscopy (TPRS) of ethanol on MgSi-

WK

To further understand the reaction mechanism, temperature-programmed reaction

spectroscopy (TPRS) was carried out using the ethanol reactant and intermediate reactive

Figure 5.5. Left: Comparison of OH groups of MgSi-WK and MgSi-IWI as probed by

in-situ dehydrated DRIFTS experiments; right: OH groups of WK catalysts with

different oxide ratios.

Page 188: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

168

molecules. The TPRS was performed on MgSi-WK surface due to the higher reactivity

than MgSi-IWI. Surface chemistry of this catalyst was previously interrogated using in

situ DRIFTS36 with the reaction mechanisms proposed based on the studied surface species

formed during the reaction. TPRS provides vapor-phase analysis allowing to fully

understand the mechanism as a prelude to the analysis of the catalytic sites responsible.

Table 5.2 shows the m/z utilized for reactive vapor (gas) species detection.

Table 5.2 m/z selection to identify the arising vapor-phase species from TPRS experiments

m/z Species

46 ethanol

26 ethylene

44 acetaldehyde

2 hydrogen

54 1,3-BD

70 crotonaldehyde

57 crotyl alcohol

MgSi-WK catalyst was pretreated at 500 °C for 1 hour to simulate the real operating

conditions. The adsorption of ethanol was performed at 100 °C to avoid any residual water

vapor condensation and flushed with inert gas to remove loosely bound molecules.

Experiment was performed under constant ethanol feed flow and shown in Figure 5.6

(left). The ethanol signal continuously decreased during the reaction as a function of

temperature without the presence of any products detected. This is consistent with the

previous report where significant amount of reactive intermediates was bound strongly to

the catalyst suface.36 Ethylene was the first product to be detected at ~200 oC which can be

explained by ethylene’s lower desorption energy than acetaldehyde.16 Acetaldehyde, when

formed, would tend to stay on the surface to undergo several other surface reactions, such

as aldol condensation and polymerization.36 The more significant consumption of ethanol

Page 189: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

169

taking place at 300 oC, where acetaldehyde and hydrogen were detected at the same time

around 350°C, which can be explained by the accelerated dehydrogenation reaction. Very

low signal of crotyl alcohol and crotonaldehyde were also evident from the spectra which

indicated the tendency of these species to undergo surface reaction than desorb off the

surface.

Evolution of the reactive intermediates and byproducts during acetaldehyde

temperature programmed experiment is shown in Figure 5.6 (right). When only

acetaldehyde was reacted over the sample crotonaldehyde was formed, in good agreement

with DRIFTS experiments reported previously.36 Several desorption temperature peaks

were observed at 210, 330, 410 °C (acetaldehyde) and 210, 350, 422 °C (crotonaldehyde).

The low temperature peak, i.e. 210 °C was due to the low temperature aldol condensation

between the two acetaldehyde molecules. The presence of ethanol during the experiment

imposes competitive surface MPV reduction and desorption of crotonaldehyde, as

previously suggested36 and in the absence of ethanol led to higher desorption rate of

crotonaldehyde. The presence of crotonaldehyde was supported by the presence of m/z=45.

Figure 5.6. TPRS spectra of ETB reaction over MgSi-WK with ethanol as the feed (left)

and acetaldehyde as the feed (right). EtOH: ethanol; AA: acetaldehyde; CA:

crotonaldehyde; C-OH: crotyl alcohol.

Page 190: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

170

This m/z can also be associated to 3-hydroxybutanal (acetaldol). The spectrum looks very

similar to that of crotonaldehyde which indicates a facile dehydration of the acetaldol

formed on the surface during the aldol condensation. Reverse reaction of acetaldol on the

surface was also expected when ethanol was not present on the surface since it would shift

the equilibrium to the left when the resulting crotonaldehyde is not reacted.49 This is

suggested by the relatively lower intensity of the m/z=45 between 300-350 °C than that of

crotonaldehyde in combination with the increasing acetaldehyde signal. The sudden

change in the slope of the TP (temperature-programmed) peak of all discussed m/z, i.e. 44,

45, and 70, indicates an additional different reaction mechanism for aldol condensation.

Palagin et al. suggested an alternative mechanism for aldol condensation without the

enolization step.50 Comparison between H-D exchange experiments of Sn-BEA, Zr-BEA

and Ti-BEA demonstrated that enolized acetaldehyde was only stabilized over Sn-BEA. A

separate mechanism took place where an open Lewis-bound acetaldehyde interacted with

a second acetaldehyde adsorbed on the opposing OH group of the catalyst. DFT calculation

showed that the activation energy of this second mechanism was more than triple than that

of the enolization mechanism (~2 eV v ~0.6 eV).50 A second peak at 330 and 350 °C for

acetaldehyde and crotonaldehyde, respectively, indicated a secondary reaction that takes

place. As previously suggested,36,51 accumulation of acetaldehyde on the catalyst surface

will lead to aldol condensation between crotonaldehyde and acetaldehyde to yield 2,4-

hexadienal. This reaction was confirmed by the change of slope from the water signal

m/z=18 which increased when the other signals decreased.

Finally, TPRS experiments were conducted with both ethanol and acetaldehyde

(Figure 5.7). Acetaldehyde was first preadsorbed on the surface, flushed with an inert gas

Page 191: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

171

to remove the physisorbed molecules and temperature ramp was performed under a

constant ethanol flow. This experiment mimics a two-step reactive process where

acetaldehyde is cofed with ethanol. If acetaldehyde production is the rate-determining step

(RDS) any accumulation of acetaldehyde on the surface would increase 1,3-BD production.

Surprisingly, this experiment did not improve production of 1,3-BD as one would expect

ethanol immediately undergo MPV reduction with the produced crotonaldehyde on the

surface. Rather, 1,3-BD production was low until 360 °C which is much later for the

ethanol alone. The sudden increase of the 1,3-BD production was accompanied by water

production suggesting the dehydration of crotyl alcohol was lagging until 360 °C as well.

This production onset coincided with a marked ethanol signal decline suggesting the MPV

reduction becoming the RDS then acetaldehyde/crotonaldehyde is accumulated on the

surface. The increase of acetaldehyde signal was mostly from both activated ethanol

dehydrogenation and MPV reduction byproduct since H2 also increased at higher

temperatures.

Figure 5.7. TPRS spectra of ETB reaction over MgSi-WK with ethanol and

acetaldehyde as the coreactants. Acetaldehyde is pre-adsorbed on the surface and

temperature ramp is under ethanol flow.

Page 192: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

172

2.4. Acid-base characterization using DRIFTS

The acidity and basicity of the catalysts was investigated using CO2 and pyridine as

probe molecules (Figure 5.8). DRIFTS was used to qualitatively describe the acid and

basic sites that are present on the surface. CO2 adsorbs on basic surface sites as surface

carbonate and bicarbonate species.52 The formation of these surface species can be

associated with the strength of the corresponding basic sites.34,36 Adsorption of CO2 on

MgSi-WK resulted in three major, broad peaks at 1655, 1541 and 1406 cm-1, while MgSi-

IWI showed broad peaks at 1645, 1620, 1505, 1406 and 1385 cm-1. These peaks were

assigned to carbonates and bicarbonate formation on MgO site since CO2 adsorption on

SiO2 should not yield any surface species.3 In particular, the fundamental doubly

degenerate υ3(COO-) vibration of carbonate on MgO is assigned to bidentate with ~1650

(υ3as) and ~1300 cm-1 (υ3

s) and monodentate carbonate ~1550 (υ3as) and ~1400 cm-1 (υ3

s)

while bicarbonate is detected at ~1650 (υ3as) and ~1380 cm-1 (υ3

s).53,54 These three species

were present on MgSi-WK catalysts, demonstrated by peaks at 1655 and 1325 cm-1

(bidentate carbonate), 1541 and ~1420 cm-1 (monodentate carbonate) and ~1670, 1406,

and 1220 cm-1 (bicarbonate). On MgSi-WK, the dominant peaks were those originating

from monodentate carbonate and bicarbonate. MgSi-IWI, on the other hand, exhibited

rather different surface chemistry. Monodentate carbonate was apparent from the peaks at

1505 and 1385 cm-1, while peaks at 1645 and 1406 cm-1 were assigned to adsorbed

bicarbonate. The presence of bidentate carbonate might be signified by the peak at 1620

and a very small peak at ~1300 cm-1. The strength of the same species on these two catalysts

was very different, as shown by the corresponding DRIFT spectra acquired at 450 °C. At

this reaction temperature, peaks at 1561 and 1368 cm-1 were apparent either due to the

Page 193: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

173

surface species rearrangement or due to the new species of lower intensity not detected at

lower temperatures. These two peaks were assigned to monodentate carbonate due to the

narrow υ3 split and stability at higher temperature.34 The stronger basicity of the MgSi-IWI

originated from its Mg-rich surface, as supported by the LEIS spectra.

Pyridine is a weak base and can discriminate between strong and basic site although

its use is limited by its relatively large molecule size, in comparison to NH3.55 Pyridine

adsorbs on a catalyst surface as a physisorbed molecule, Lewis-bonded species and as a

pyridinium ions.55–57 Observation of the latter two species allows to discriminate between

the LAS and Brønsted acid sites (BAS).58 BAS, which would be indicated by a peak at

1638 and 1540 cm-1,58 was not observed which aligns well with the observations of

Angelici et al. and Janssens et al.9,30 Both catalysts exhibited similar Lewis acidity as

shown at 450 °C as evident from peaks at 1445, 1590 and 1605 cm-1. The peaks at 1445

and 1608 cm-1 shifted to 1450 and 1608 cm-1 at higher temperature, i.e. 450 °C, while the

peak at 1490 cm-1, which is a combination band of both LAS and BAS, disappeared. Based

on both HS-LEIS and DRIFTS experiments, the origin of this Lewis acidity should be

different for both catalysts where MgSi-WK would acquire its acidity from unpaired

Figure 5.8. Acid-base characterization of MgSi-IWI and MgSi-WK catalysts probed

using CO2 (left) and pyridine (right). Spectra at high temperature (450°C) and low

temperature (100°C) are shown.

Page 194: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

174

electrons of oxygen atom in the Mg-O-Si coordination while MgSi-IWI mostly from the

Mg-O-Mg coordination (vide supra).

2.5. Reactive site persistence during ethanol-to-1,3-BD

Far less explored is reactive site acidity and basicity characterization (and

persistence of the active sites) under reactive conditions of ethanol at reaction temperatures.

To elucidate the role of the specific basic and acidic sites during the reaction on the aged

catalyst, in-situ characterization experiments were performed using CO2 and pyridine after

ethanol reaction on MgSi-WK. In the first experiment ethanol was adsorbed at 100 °C for

20 min, flushed with inert gas for 1 hour and the resulting surface sites were probed with

CO2 or pyridine in-situ. In the second experiment, reaction with ethanol was carried out at

200 °C to initiate ethanol dehydrogenation. After 1 hour of reaction the reaction cell was

flushed and CO2 or pyridine was introduced into the cell. This way nature of the reactive

sites, their persistence and availability for reaction were measured.

Figure 5.9 (left) shows spectra resulting from CO2 adsorption. Three distinct peaks

at 1615, 1380 and 1330 cm-1 appeared on the catalyst that was previously exposed to

Figure 5.9. In-situ acid-base characterization of MgSi-WK catalyst before and after

ethanol adsorption at 100°C and reaction at 200°C using CO2 (left) and pyridine (right).

Page 195: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

175

ethanol (MgSi-WK EtOH(ads)) at 100 °C that do not originate from CO2. While peaks at

1380 and 1330 cm-1 were assigned to ρw (CH) and ρw (CH3) of surface ethoxide,36 peak at

1615 cm-1 can be associated with a monodentate carbonate, accompanied by the

symmetrical vibration peak at ~1320 cm-1. After ethanol adsorption all peaks decreased in

intensity, as compared to CO2-only adsorption on the unreacted catalyst at the same

temperature. This significant decrease indicates a competitive adsorption between ethanol

and CO2 that resulted in surface ethoxide and both monodentate and bidentate carbonate.

This indicates that ethanol preferably adsorbed on the strong basic sites that would form

monodentate and bidentate carbonate when exposed to CO2. Reaction at 200 °C was

performed in-situ before degassing with inert and adsorption of CO2. The reaction

temperature was chosen of 200 °C to limit further reactions to 1,3-BD. Extensive inert

degassing was done to limit the further C4 oxygenates formed from facile aldol

condensation, dehydration, and polymerization from occupying the site.36 Comparison of

the spectra between CO2 adsorbed on activated catalyst and aged catalyst, i.e. extensively

reacted at 200°C, showed decrease in intensity on all peaks related to all surface species

arising from CO2. Weak basic sites, which are represented by peaks at ~1650 and ~1400

cm-1 (surface bicarbonate),53 were depleted. This indicates the consumption of the weak

basic sites during the reaction. These weak basic sites on MgO/SiO2, would be the OH

groups, since the strong Mg2+-O2- pairs form monodentate and bidentate carbonate when

exposed to CO2.

Pyridine probing of the acidic sites shows a non-discriminative trend for both

ethanol adsorption at 100 oC and reaction at 200 oC. LAS indeed were consumed during

the adsorption and even more so after the reaction at 200°C, as indicated in Figure 5.9

Page 196: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

176

(right). No generation the new of BAS was observed after either adsorption or reaction as

indicated by the absence of peaks at 1545 and 1638 cm-1.58 The Mg-O-Mg linkage present

on MgO exhibits Lewis acidity to a certain extent, with the absence of Brønsted acidity.50

The LAS on MgSi-WK are fundamentally represented by the Mg atoms in four groups:

Mg-O(H)-Mg, Mg-O(H)-Si, Mg-O-Mg, and Mg-O-Si, with the first two groups being the

open LAS. These were further distinguished by coordination number of the magnesium

atoms: the lower the coordination, the stronger the atom is due to the electron deficiency

of the cation. Although pyridine can’t discriminate open LAS from closed sites, its

combination with the conducted CO2 DRIFTS experiment provides more information on

the involved group. The two sites are distinguished by the consumed bicarbonate species,

which is formed when CO2 is adsorbed to a site containing OH group. Hence, the

consumption of both LAS and bicarbonate site (weak basic site), can be traced to the open

LAS. The two open sites, i.e. Mg-O(H)-Mg and Mg-O(H)-Si, were discriminated by the

strength of the base pair. The former is less likely to participate during the reaction, due to

its very basic nature. However, activation of this group has been observed when a bare

MgO is activated using NH3-thermally treated MgO.12 The Mg-O-Si linkages have

previously been correlated to the enhanced activity1,2 while open LAS had been shown to

be responsible for the increased 1,3-BD production.19

To further elucidate the role of acidic and basic sites during the reaction to 1,3-

butadiene, surface site poisoning experiments were carried out using probe molecules such

as CO2, propionic acid and NH3 in a steady state fixed bed reactor. CO2 and propionic acid

are two weak acids while NH3 is a basic probe molecule. Figure 5.10a shows the effect of

cofeeding with CO2. Slow, steady decrease in acetaldehyde and 1,3-BD production before

Page 197: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

177

CO2 was introduced indicates slow catalyst deactivation. However, once the CO2 was cofed

to the system, formation rate of 1,3-BD and ethylene dramatically dropped, while

acetaldehyde production increased. CO2 is a weaker acid than propionic acid and will bind

to the strongest basic sites. As shown in Figure 5.8, CO2 did not typically adsorb to the

surface at 450 °C. This poisoning effect suggests that CO2 poisoned the sites that catalyzed

aldol condensation and the subsequent steps, since more acetaldehyde was released into

the vapor-phase without further reacting. When CO2 flow was switched off, the production

of acetaldehyde, 1,3-BD and ethylene was restored, confirming the weak interaction

between CO2 and the strong basic sites.

Figure 5.10b shows (by)product formation rates upon the introduction of propionic

acid. All three products showed a decline in formation rate. When propionic acid

concurrent flow was stopped, the production of acetaldehyde was restored but 1,3-BD and

ethylene formation did not recover. Propionic acid interacted more strongly with the

stronger base sites but also binds to any weaker basic sites. Hence, when propionic acid

flow was stopped, only weak basic sites were accessible while some of the strong basic

sites were permanently poisoned. From the two experiments it is evident that acetaldehyde

production was catalyzed by weak basic sites and 1,3-BD - by strong basic sites. The

production of acetaldehyde over the weak basic site is consistent with the DRIFTS

performed using CO2 as probe molecule (Figure 5.9). Similar phenomenon was observed

by Shylesh, et al. where the 1,3-BD formation rate did not recover during propionic acid

cofeeding experiment over the Au-promoted IWI MgO/SiO2 catalyst.3 Very interestingly,

ethylene formation during the cofeeding experiments followed the trend of 1,3-butadiene.

As previously suggested, ethylene formation can be carried out over both Lewis acidic

Page 198: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

178

oxygen atoms (LAS) in the Mg-O-Mg or Mg-O-Si sites and the acidic O-H group

(BAS).16,59 Poisoning with both CO2 and propionic acid affected the strong and medium

Lewis basic Mg atoms in the Mg-O-Mg and Mg-O-Si which inevitably perturbed the strong

Lewis acid pair, i.e. oxygen anions, as well.

Figure 5.10. Acid-base poisoning reactivity testing using (a) CO2, (b) propionic acid,

and (c) NH3 to determine the role of each site during ethanol conversion to 1,3-BD over

WK-800 MgO/SiO2 catalyst. Reactions are carried out at 425 °C, mcat = 0.1 g, pethanol =

2.5 kPa, total flow = 55 ml/min. All formation rates are normalized to initial 1,3-BD

formation rate.

Page 199: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

179

NH3 is a relatively strong gas-phase base, stronger than pyridine and other organic

basic molecules, such as acetonitrile and benzenes.55 At reaction temperature of 425 °C,

NH3 exhibits very weak adsorption on the surface and would interact with the acidic sites.

Evident from Figure 5.10c, acetaldehyde production was hardly affected by the poisoning,

as opposed to 1,3-BD and ethylene, where the decrease in production was very pronounced.

While ethylene formation inhibition was reversible, 1,3-BD formation was irreversibly

affected. NH3 poisoned both strong and weak BAS and LAS but when its flow was

discontinued only the strong BAS were poisoned. Ethylene synthesis trend was very similar

for both propionic acid and NH3 cofeeding which indicates the same acid-base pairs being

poisoned during the experiment. 1,3-BD production involves two dehydration steps and

the poisoning indicates that its production did not require participation from the site that

dehydrates ethylene. From these experiments, it is evident that dehydration steps of both

acetaldol and crotyl alcohol were catalyzed by strong acidic sites while ethanol dehydration

was catalyzed by weaker acidic sites.

To confirm participation of the acidic sites, Na2O, a strong basic oxide, was added

to the catalyst in a post-treatment step that permanently poisoned some of the acidic sites.

Three Na concentrations were chosen in this study, i.e. 250, 500, and 1000 ppm, labeled as

250NaMgSi-WK, 500NaMgSi-WK, and 1000NaMgSi-WK. Ethylene productivity was

significantly limited with up to 50% suppression, as shown in Figure 5.11b. Figure 5.11a

shows that sodium poisoning limited the productivity of 1,3-BD as well. The only condition

that gave comparable 1,3-BD productivity to the undoped catalyst was at 450 °C and with

250NaMgSi-WK catalyst. This limitation was also reflected in the acetaldehyde

production. As in 1,3-BD case there was no trend observed with varying Na content but

Page 200: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

180

acetaldehyde production over Na2O-poisoned catalysts was higher than for the unpromoted

catalysts. This suggests that Na2O poisoned the acidic sites unselectively eliminating both

BAS and LAS that are responsible for ethylene and 1,3-BD production. The effect of

acidity modification by poisoning with alkali metal was also investigated by Da Ros et al.60

However, the loss of acidic sites in their catalysts was offset by the presence of Zr and Zn

which provided extra Lewis acidity, which was presumably responsible for selectively

dehydrate the C4 molecules to 1,3-BD.

The origin of the poisoning effect was correlated with CO2 and NH3 DRIFTS

experiments. Figure 5.12a shows the appearance of the peaks at 1643, 1463, 1372, and

1356 cm-1 in addition to the native basic sites of the MgSi-WK. Peaks at 1463 and 1356

cm-1 reached a maximum for 500NaMgSi-WK before decreasing in intensity at 1000 ppm,

while the shoulder at 1372 cm-1 became obvious with higher Na loading. The stability of

these peaks indicates the presence of bicarbonates, as shown by the peaks at 1643 and 1372

(1356) cm-1. The peak at 1463 cm-1 is sharp and also is not stable at high temperature which

indicates the presence of another weakly adsorbed CO2 species. Na2CO3, on the other hand,

possesses a band at this specific wavenumber, i.e. 1463 cm-1.61 Na2CO3 is a very stable

carbonate melting prior to decomposing to Na2O, starting at 850°C.62 At higher

temperature range, i.e. 300-450°C an interesting observation emerges from the growth of

the peaks at ~1600 and ~1370 cm-1, as shown in Figure 5.12a (top) for 1000NaMgSi-WK.

The presence of the shoulders to these peaks suggests the presence of two distinct carbonate

groups. From the split, i.e. ~300 cm-1, these peaks are characteristics of bidentate carbonate.

The growth of this peak suggests that rearrangement of (bi)carbonates took place upon

Page 201: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

181

thermal treatment. Furthermore, the stability of this peak at high temperature indicates the

enhanced basic site when Na was introduced to the catalyst.

The acidity of the Na-modified catalysts was probed with using NH3 as a probe

molecule, as shown in Figure 5.12b. On pure dehydroxylated MgO, NH3 adsorbs as

physisorbed molecule, with a peak at ~1605 cm-1 while two different LAS exhibited

vibrational peaks at >1605 and 1560 cm-1.63,64 On hydroxylated surface, however, the

interaction is more complicated due to the contributions of OH groups, where hydrogen

Figure 5.11. Productivity of (a) 1,3-BD, (b) ethylene, and (c) acetaldehyde of Na-

poisoned MgSi-WK catalysts between 350-450°C. Catalyst mass = 0.1 gr, total flow rate

= 55 ml/min, pethanol = 2.5 kPa.

Page 202: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

182

bonding between NH2H—OH (1612 cm-1) and H3N—HO (1634 cm-1) obscured the

DRIFTS spectra.63,64 The peak at 1430 cm-1 was assigned to ammonium ion (NH4+) as a

result from interaction between ammonia and a BAS.63 Physisorbed ammonia species was

observed on both oxygen and hydroxyl group by the appearance of peaks at 1554 and 1605

cm-1, respectively.63,64 Presence of BAS during NH3 but not pyridine probing had been

observed in the past, where pyridine underrepresented the amount of acidic sites.9,65 This

discrepancy was due to the size of the molecule with NH3 being more mobile.9 The BAS

that was found on the catalysts was isolated, less-accessible and hence might participate

less during the reaction. Four peaks that are assignable to LAS are recognized on these

catalysts by the peaks at 1560, 1580, 1600, 1620, and 1650 cm-1. To aid the assignments

of these peaks DFT vibrational frequency calculations performed on defect sites of MgO,

i.e. Mg3CO4C and Mg4CO4C for both open and closed LAS (Figure 5.13). The corresponding

vibrations were tabulated in Table 5.3.

Table 5.3. Comparison between observed experimental values of NH3 adsorption on MgSi-

WK catalysts with DFT calculated IR vibrations of NH3 adsorbed on open and closed acid

Mg3C and Mg4C sites. Scaling factor of 0.9854 was applied to the calculated values and was

derived from the gas-phase NH3 experimental and DFT calculated frequencies.

Type Mg coordination Vibrational mode Experimental values

δas H-N-H δs H-N-H δas H-N-H δs H-N-H

Open 4C 1592 1566 1600 1560

3C 1574 1534 1570 1540

Closed 4C 1588 1571 N/A N/A

3C 1620 1577 1620 1580

Assuming an ideal surface and similar trend on Mg-O-Si sites, peak assignments

can be readily made. For instance, experimentally, MgSi-WK (0 ppm Na) catalyst

exhibited both closed and open LAS. Closed LAS were evident from the peaks at 1620 and

1580 cm-1 (Mg3C) while open LAS from Mg4C were recognized by peaks at 1600 and 1560

Page 203: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

183

cm-1. The presence of peak at 1540 cm-1 signified the presence of another open LAS (Mg3C)

which would be accompanied by a shoulder peak at ~1570 cm-1. The assignments of these

peaks here were made based on the comparison with the DFT computed values tabulated

in Table 5.3. The shoulder at 1650 cm-1 is obvious and can’t be ignored. Echterhoff and

Knözinger attributed a peak at 1634 cm-1 to hydrogen bonding between ammonia and

surface hydroxyl (NH2-H—HO-Mg).64 This vibrational mode is entirely possible due to

the identified NH4+ on the surface indicating the presence of some BAS. Alternatively, if

a Si atom replaces one Mg atom in the Mg3C-O4C-Mg4C (closed LAS) and results in Mg-

O-Si linkage then a change in electronegativity will occur and the magnesium atom

becomes more positively charged which would result in the shorter bond between the Mg

and N atoms. The shorter bond would result in the shift of the peak to a higher wavenumber

in this case higher than 1620 cm-1. Increasing Na loading resulted in the decreasing LAS

and BAS, which was expected. The 250 ppm Na-poisoned catalyst, surprisingly, exhibited

higher intensity of all peaks associated with LAS and BAS, indicating the enhanced acidity

at this temperature. The enhanced acidity, however, was only intermittent at this

temperature, since the top spectra in Figure 5.12b shows the better stability of NH3 surface

species on unpoisoned catalyst at elevated temperature, i.e. 300 °C.

Page 204: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

184

Figure 5.12. Bottom: DRIFTS characterization of Na-doped MgSi-WK using (a) CO2

and (b) NH3. Spectra are taken at 100°C after extensive evacuation with N2. Top: (a)

CO2 desorption spectra of 1000 ppm Na-doped MgSi-WK at 100, 300, and 450°C and

(b) NH3 desorption on 0 ppm and 250 ppm Na-doped MgSi-WK at 300°C. Spectral

subtraction was done using the spectra of the dehydrated catalysts at respective

temperatures as the background.

Page 205: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

185

(b)

Figure 5.13. (a) MgO periodic model used for DFT simulation of NH3 adsorption on

MgO Lewis acid sites: (b) Mg3C, closed, (c) Mg4C, closed, (d) Mg3C, open, (e) Mg4C,

open. Multiple possible adsorption sites, i.e. kink (Mg3CO4C), edge (Mg4CO4C), and

planar (Mg5CO5C) are highlighted.

(a)

(c)

(d) (e)

Mg3C

Mg3C O4C

O4C

Mg4C

Mg4C

O4C

O4C

Mg4C

O4C

O4C

O4C Mg4C

Mg4C

O5C

Mg3C O4C Mg4

C

O4C

Mg5C O5C

Page 206: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

186

2.6. Implications for the structure-activity relationship

The wet-kneading method provides a deeper, more intimate mixing between MgO

and SiO2.1 Incipient wetness impregnation, on the other hand, is a typical synthesis method

for the synthesis of supported catalyst. This method is appropriate when the target catalyst

is a well-dispersed, below monolayer, metal oxide that is supported on a high surface area

support. However, at ratio of 1, the ‘promoter’ is at similar amount with the support, which

corresponds to layered metal oxide on the surface. The nature of the synthesis method

would result in two different bulk phases, i.e. MgO-phase and SiO2-phase, which are

bridged by an interface that should equally contain both oxides. Figure 5.14 schematically

represents these three phases that are formed during the synthesis method. In the case of

wet-kneading, the extensive interaction between MgO and SiO2 allows the boundary phase,

i.e. the middle part in Figure 5.14, to grow larger, as confirmed by the LEIS experiment

Figure 5.14. Schematic diagram to show the presence of various sites investigated with

NH3 and CO2 DRIFTS experiments. The basic sites (orange) are shown in the figure as

both Brønsted base (OH) and Lewis site (electron accepting oxygen atoms), and acid sites

(blue) are represented as Brønsted acid sites (H) and Lewis acid sites (electron donating

magnesium and silicon atoms).

Page 207: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

187

and in-situ dehydrated DRIFTS of the catalysts. The MgO-rich phase that is formed on the

catalyst contributes to the higher amount of strong basic sites, which is evident from CO2-

probing comparison, shown in Figure 5.8. The reduced amount of strong basic sites makes

MgSi-WK a better catalyst than MgSi-IWI, as demonstrated by the higher 1,3-BD

formation.

As confirmed by TPRS, the catalytically relevant step during the transformation of

ethanol to 1,3-BD was determined to be acetaldehyde formation step. The ex-situ

characterization alone indicated the importance of weaker basic sites during the reaction

and in-situ poisoning experiment with CO2 and propionic acid confirmed the need for the

weaker basic sites for the reaction. In particular, poisoning the strong basic sites with CO2

resulted in higher acetaldehyde formation rate which suggests that the weak basic sites, not

poisoned by CO2, catalyzed ethanol dehydrogenation and the stronger basic sites catalyze

the subsequent reaction steps. This finding agrees that of Shylesh et al. where propionic

acid was used as a poisoning agent.3 The acidic sites are responsible for both 1,3-BD and

ethylene formation considering the nature of the reaction steps. From NH3 poisoning

experiments it is evident that the weaker acidic sites are responsible for ethanol dehydration

while the stronger sites are responsible for acetaldol and crotyl alcohol dehydration. The

reasoning behind this is the strong interaction between crotyl alcohol, crotonaldehyde and

the surface.36 Poisoning the strong acidic sites resulted in the accumulation of the heavy

aromatic compounds on the surface, which is also confirmed by the unprecedented amount

of aromatic carbonaceous compound observed after the NH3 reaction. The indispensable

role of acidic sites on the reaction is further confirmed by Na-poisoned catalysts, where,

Page 208: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

188

although ethylene formation was virtually suppressed, 1,3-BD formation did not

significantly benefit from the reduced acidic sites.

As suggested by the characterization using in-situ DRIFTS and LEIS, the mixing

between MgO and SiO2 would allow the formation of Mg-O-Si linkages which is

consistently observed by previous investigators.1–3 With the possibility of hydroxylation of

the surface, there are four possible formation of the LASs, i.e. open and closed acid sites

of both Mg-O-Mg and Mg-O-Si. The LASs, i.e. Mg2+ cations, will present different

strength, depending on the coordination and type of linkage. Our NH3-DRIFTS experiment

demonstrated the formation of several distinct Lewis sites of different strength. The in-

situ acid-base characterization after ethanol reaction at 200 °C showed that weak basic sites

and LASs were depleted. Schematic representation of this site can be seen in Figure 5.15.

Figure 5.15. Representation of the role of basic sites during ethanol conversion to

acetaldehyde. Top figure represents dehydrated (pretreated) catalyst; bottom figure

demonstrates the absence of bicarbonate when CO2 is adsorbed in-situ after reaction at

200 °C.

Page 209: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

189

The weak basic sites discussed are both the Mg-O(H)-Mg and Mg-O(H)-Si. Since at 200

°C only ethylene and acetaldehyde were produced this suggested that the open sites were

the most catalytically relevant active sites. It should be noted, however, that this linkage

contains both acidic and basic sites in the form of Mg2+ cations and OH group, respectively.

Hence, the strength of Mg-O(H)-Mg and Mg-O(H)-Si should be different. The basicity for

the latter would be weaker, as suggested by the in-situ DRIFTS, where the peaks for these

linkages are well below that of Mg-OH, i.e. 3680, 3705, and 3725 cm-1 (Figure 5.5). The

strength of the acidic site, i.e. Mg cation, can be hypothesized to be lower as well in the

case of Mg-O(H)-Si due to the electron density, since Si cations possess atomic charge of

+4e. This leads to the proposed both Mg-O(H)-Si to be the sites that are responsible for

ethanol dehydrogenation and ethylene dehydration while stronger acidic sites and basic

sites are responsible for C4 dehydration and ethanol dehydrogenation. The exact molecular

structure of the various open (closed) acidic and basic sites proposed is shown in Figure

5.14 and includes both LAS and BAS.

3. Conclusions

MgO/SiO2 catalyst active surface sites were analyzed using in situ DRIFTS (using

complementary DFT calculations), TPRS and steady state reactor in combination with bulk

XRD and surface LEIS measurements. Acid-base characterization showed that IWI

synthesis method resulted in a highly basic catalyst with reactive properties originating

from the abundance of Mg atoms on the topmost surface layer, as opposed to WK catalyst.

The molecular active site structure was determined and MgSi-IWI surface was found to be

dominated with stabilized Mg-OH with little magnesium silicate hydroxyl groups. The

Page 210: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

190

MgSi-WK surface, on the other hand, contained a significant number of surface sites

derived from magnesium silicates as indicated by the distinct OH groups. This fundamental

site structure difference consequentially led to a different reactivity where MgSi-WK

possessed a more balanced weak-strong basic sites than the basic sites present on MgSi-

IWI. From various reacting molecule poisoning experiments it was determined that the

weak basic sites were responsible for ethanol dehydrogenation, strong basic sites for aldol

condensation and MPV reduction, while stronger acid sites catalyze acetaldol and crotyl

alcohol dehydration reactions and weak acid sites catalyzed the undesired ethanol

dehydration. Furthermore, through a combination of NH3-TPD and DFT the presence of

open and closed LAS was identified while further elaborating Mg coordination, as adopted

from LAS classification of zeolitic materials.20–22 The MgSi-WK catalyst was shown to

have both open LAS with both Mg3C and Mg4C as the anchoring LAS, while also a very

isolated closed LAS (Mg3C).

Page 211: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

191

References

(1) Chung, S.-H.; Angelici, C.; Hinterding, S. O. M.; Weingarth, M.; Baldus, M.;

Houben, K.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ACS Catal. 2016, 6 (6),

4034–4045.

(2) Velasquez Ochoa, J.; Bandinelli, C.; Vozniuk, O.; Chieregato, A.; Malmusi, A.;

Recchi, C.; Cavani, F. Green Chem. 2016, 18, 1653–1663.

(3) Shylesh, S.; Gokhale, A. A.; Scown, C. D.; Kim, D.; Ho, C. R.; Bell, A. T.

ChemSusChem 2016, 9 (12), 1462–1472.

(4) Huang, X.; Men, Y.; Wang, J.; An, W.; Wang, Y. Catal. Sci. Technol. 2017, 7 (1),

168–180.

(5) Muller, P.; Burt, S. P.; Love, A. M.; McDermott, W. P.; Wolf, P.; Hermans, I. ACS

Catal. 2016, 6 (10), 6823–6832.

(6) Sushkevich, V. L.; Ivanova, I. I.; Taarning, E. Green Chem. 2015.

(7) Shylesh, S.; Gokhale, A. A.; Ho, C. R.; Bell, A. T. Acc. Chem. Res. 2017,

acs.accounts.7b00354.

(8) Pomalaza, G.; Capron, M.; Ordomsky, V.; Dumeignil, F. Catalysts 2016, 6 (12),

203–237.

(9) Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A. Catal.

Sci. Technol. 2015.

(10) Kvisle, S.; Aguero, A.; Sneeden, R. P. A. Appl. Catal. 1988, 43 (1), 117–131.

(11) Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A.

ChemSusChem 2014, 7 (9), 2505–2515.

(12) Hayashi, Y.; Akiyama, S.; Miyaji, A.; Sekiguchi, Y.; Sakamoto, Y.; Shiga, A.;

Koyama, T.; Motokura, K.; Baba, T. Phys. Chem. Chem. Phys. 2016, 18 (36),

25191–25209.

(13) Di Valentin, C.; Del Vitto, A.; Pacchioni, G.; Abbet, S.; Wörz, A. S. S.; Judai, K.;

Heiz, U. J. Phys. Chem. B 2002, 106 (46), 11961–11969.

(14) Cavalleri, M.; Pelmenschikov, A.; Morosi, G.; Gamba, A.; Coluccia, S.; Martra, G.

In Oxide-based Systems at the Crossroads of ChemistrySecond International

Workshop October 8-11, 2000, Como, Italy; A. Gamba, C. C. and S. C. B. T.-S. in

S. S. and C., Ed.; Elsevier, 2001; Vol. Volume 140, pp 131–139.

(15) Branda, M. M.; Rodríguez, A. H.; Belelli, P. G.; Castellani, N. J. Surf. Sci. 2009,

603 (8), 1093–1098.

(16) Taifan, W. E.; Bučko, T.; Baltrusaitis, J. J. Catal. 2017, 346, 78–91.

(17) Chieregato, A.; Velasquez Ochoa, J.; Bandinelli, C.; Fornasari, G.; Cavani, F.;

Mella, M. ChemSusChem 2014, 8 (2), 377–388.

(18) Fan, D.; Dong, X.; Yu, Y.; Zhang, M. Phys. Chem. Chem. Phys. 2017, 19 (37),

25671–25682.

(19) Sushkevich, V. L.; Palagin, D.; Ivanova, I. I. ACS Catal. 2015, 4833–4836.

(20) Harris, J. W.; Cordon, M. J.; Di Iorio, J. R.; Vega-Vila, J. C.; Ribeiro, F. H.;

Gounder, R. J. Catal. 2016, 335, 141–154.

(21) Boronat, M.; Concepcion, P.; Corma, A.; Navarro, M. T.; Renz, M.; Valencia, S.

Phys. Chem. Chem. Phys. 2009, 11 (16), 2876–2884.

(22) Boronat, M.; Concepción, P.; Corma, A.; Renz, M.; Valencia, S. J. Catal. 2005,

234 (1), 111–118.

(23) Kantorovich, L. N.; Holender, J. M.; Gillan, M. J. Surf. Sci. 1995, 343 (3), 221–

Page 212: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

192

239.

(24) Chizallet, C.; Petitjean, H.; Costentin, G.; Lauron-Pernot, H.; Maquet, J.;

Bonhomme, C.; Che, M. J. Catal. 2009, 268 (1), 175–179.

(25) Chizallet, C.; Costentin, G.; Che, M.; Delbecq, F.; Sautet, P.; Surface, D.; Marie,

P. J. Am. Chem. Soc. 2007, No. 19, 6442–6452.

(26) Chizallet, C.; Costentin, G.; Che, M.; Delbecq, F.; Sautet, P. J. Phys. Chem. B

2006, 110 (32), 15878–15886.

(27) Sushko, P. V.; Shluger, A. L.; Catlow, C. R. A. Surf. Sci. 2000, 450 (3), 153–170.

(28) Bailly, M.; Chizallet, C.; Costentin, G.; Krafft, J.; Lauronpernot, H.; Che, M. J.

Catal. 2005, 235 (2), 413–422.

(29) Sushkevich, V. L.; Vimont, A.; Travert, A.; Ivanova, I. I. J. Phys. Chem. C 2015,

119 (31), 17633–17639.

(30) Janssens, W.; Makshina, E. V. V.; Vanelderen, P.; De Clippel, F.; Houthoofd, K.;

Kerkhofs, S.; Martens, J. A. A.; Jacobs, P. A. A.; Sels, B. F. B. F. ChemSusChem

2014, 8 (6), 994–1008.

(31) Larina, O. V.; Kyriienko, P. I.; Trachevskii, V. V.; Vlasenko, N. V.; Soloviev, S.

O. Theor. Exp. Chem. 2016, 51 (6), 387–393.

(32) Zhu, Q.; Wang, B.; Tan, T. ACS Sustain. Chem. Eng. 2017, 5 (1), 722–733.

(33) Golay, S.; Kiwi-Minsker, L.; Doepper, R.; Renken, A. Chem. Eng. Sci. 1999, 54

(15), 3593–3598.

(34) Di Cosimo, J. I.; Dıez, V. K.; Xu, M.; Iglesia, E.; Apesteguıa, C. R. J. Catal. 1998,

178 (2), 499–510.

(35) Phung, T. K.; Busca, G. Catal. Commun. 2015, 68, 110–115.

(36) Taifan, W.; Yan, G. X.; Baltrusaitis, J. Catal. Sci. Technol. 2017, 7 (20), 4648–

4668.

(37) Gao, M.; Jiang, H.; Zhang, M. Appl. Surf. Sci. 2018.

(38) Zhu, Q.; Wang, B.; Tan, T. ACS Sustain. Chem. Eng. 2017, 5 (1), 722–733.

(39) Kotov, S. V.; Nankeva, I. N. Chem. Technol. Fuels Oils 1994, 30 (5), 240–245.

(40) Ho, C. R.; Shylesh, S.; Bell, A. T. ACS Catal. 2016, 6 (2), 939–948.

(41) Mathur, L.; Saddam Hossain, S. K.; Majhi, M. R.; Roy, P. K. Boletín la Soc.

Española Cerámica y Vidr. 2017.

(42) Tavangarian, F.; Emadi, R. Mater. Res. Bull. 2010, 45 (4), 388–391.

(43) Tavangarian, F.; Emadi, R. Ceram. – Silikáty 54 (2), 122–127.

(44) Swanson, H. E.; Tatge, E. J. Res. Natl. Bur. Stand. (1934). 1951, 46 (4), 318–327.

(45) ter Veen, H. R. J.; Kim, T.; Wachs, I. E.; Brongersma, H. H. Catal. Today 2009,

140 (3), 197–201.

(46) Davydov, A. Molecular Spectroscopy of Oxide Catalyst Surfaces; John Wiley &

Sons Ltd: West Sussex PO19 8SQ, England, 2003.

(47) Keturakis, C. Operando Molecular Spectroscopy during Catalytic Biomass

Pyrolysis, Lehigh University, 2015.

(48) Knözinger, E.; Jacob, K.-H.; Singh, S.; Hofmann, P. Surf. Sci. 1993, 290 (3), 388–

402.

(49) Toussaint, W. J.; Dunn, J. T.; Jachson, D. R. Ind. Eng. Chem. 1947, 39 (2), 120–

125.

(50) Palagin, D.; Sushkevich, V. L.; Ivanova, I. I. J. Phys. Chem. C 2016, 120 (41),

23566–23575.

Page 213: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

193

(51) Ordomsky, V. V.; Sushkevich, V. L.; Ivanova, I. I. J. Mol. Catal. A Chem. 2010,

333 (1), 85–93.

(52) Bian, S.-W.; Baltrusaitis, J.; Galhotra, P.; Grassian, V. H. J. Mater. Chem. 2010,

20 (39), 8705.

(53) Busca, G.; Lorenzelli, V. Mater. Chem. 1982, 7 (1), 89–126.

(54) Taifan, W.; Boily, J.-F.; Baltrusaitis, J. Surf. Sci. Rep. 2016, 71 (4), 595–671.

(55) Lercher, J. A.; Grijndling, C.; Eder-Mirth, G. Catal. Today 1996, 27, 353–376.

(56) Lavalley, J. C. Catal. Today 1996, 27, 377–401.

(57) Vimont, A.; Thibault-Starzyk, F.; Daturi, M. Chem. Soc. Rev. 2010, 39 (12), 4928–

4950.

(58) Busch, O. M.; Brijoux, W.; Thomson, S.; Schüth, F. J. Catal. 2004, 222 (1), 174–

179.

(59) DeWilde, J. F.; Chiang, H.; Hickman, D. A.; Ho, C. R.; Bhan, A. ACS Catal. 2013,

3 (4), 798–807.

(60) Da Ros, S.; Jones, M. D.; Mattia, D.; Pinto, J. C.; Schwaab, M.; Noronha, F. B.;

Kondrat, S. A.; Clarke, T. C.; Taylor, S. H. ChemCatChem 2016, 8, 1–12.

(61) Joshi, S.; Kalyanasundaram, S.; Balasubramanian, V. Appl. Spectrosc. 2013, 67

(8), 841–845.

(62) Kim, J.-W.; Lee, H.-G. Metall. Mater. Trans. B 2001, 32 (1), 17–24.

(63) Echterhoff, R.; Knözinger, E. J. Mol. Struct. 1988, 174, 343–350.

(64) Echterhoff, R.; Knözinger, E. Surf. Sci. 1990, 230 (1), 237–244.

(65) Angelici, C.; Meirer, F.; van der Eerden, A. M. J.; Schaink, H. L.; Goryachev, A.;

Hofmann, J. P.; Hensen, E. J. M.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ACS

Catal. 2015, 5 (10), 6005–6015.

Page 214: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

194

Chapter 6

Role of transition metal promoters (Cu, Zn)

on MgO/SiO2 catalyst for Lebedev reaction

Abstract ..........................................................................................................................194

1. Introduction .......................................................................................................195

2. Computational results ...................................................................................... 199 2.1. Model catalyst selection and analysis ...................................................... 199

2.2. Reactive intermediates ............................................................................. 206

2.3. Potential energy surfaces .......................................................................... 210

3. Experimental results .........................................................................................214 3.1. Catalyst characterization .......................................................................... 214

3.2. Steady state catalytic performance and acid/base chemistry of the

catalyst active sites .............................................................................................. 222

3.3. Active sites under operating conditions ................................................... 226

3.3.1. Temperature programmed infrared spectroscopy measurements (TP-

DRIFTS) ............................................................................................................. 226

3.3.2. In-situ UV-Vis DRS study of MgSi catalysts .......................................... 230

3.3.3. Operando XAS studies of Cu, Zn-promoted MgSi catalysts ................... 232

3.3.3.1. Operando XANES and EXAFS of Cu-promoted MgSi catalyst ......... 232

3.3.3.2. Operando XANES and EXAFS of Zn-promoted MgSi catalyst ......... 241

4. Conclusion ......................................................................................................... 245

Supporting Information ............................................................................................... 247

References ...................................................................................................................... 266

Abstract

Electronic structure and reactivity of Cu- and Zn-promoted wet kneaded MgO/SiO2

catalysts was interrogated during ethanol reaction to 1,3-BD. A multimodal nature of

characterization, including in situ or operando X-ray, electron, light spectroscopies and

steady state reactivity measurements demonstrated critical new information on the

Page 215: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

195

temporal evolution of the catalyst active sites including key measurements performed

operando using synchrotron source (EXAFS and XANES). In situ DRIFT spectroscopy

allowed to decouple the aldol condensation and dehydrogenation reactive steps due to the

promotion with enhanced ability to carry out aldol condensation, as correlated with the

steady state reactivity experiments. . In situ UV-Vis spectroscopy presented a complex

picture of the adsorbates with π- π* electronic transitions due to the allylic cations, cyclic

or aromatic species while also suggesting oligomeric CuO species were formed. Operando

X-ray measurements combined with ab initio multiple scattering modelling performed as

a function of temperature identified a new transient intermediate assigned to a 4-fold

coordinate Cu species that was key leading to increase in Cu-Cu bond number. For the

first time, two types of Zn bonds, namely Zn-O and Zn-Mg, were identified during X-ray

analysis under operating conditions. With Zn nearly 6-coordinated when in the vicinity of

Mg while Zn-O species coordinated to nearly 4 nearest neighbors. The data suggest that

such supported catalyst deactivation might proceed not only via carbon coking mechanism

but also through the dispersed Cu site diffusion and growth due to the nearest neighbor

oxygen atoms loss. The results presented suggest intermediates for

segregation/deactivation mechanisms for a broader set of supported Cu and Zn catalysts

used for alcohol upgrading catalytic reactions.

1. Introduction

Catalytic conversion of ethanol to 1,3-butadiene (1,3-BD) is a promising green and

renewable route for obtaining a commodity chemical that does not utilize a conventional

Page 216: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

196

petroleum-based feedstock.1 The feedstock and technological process landscape in 1,3-BD

production is undergoing changes due to the distinct industry shift from oil to C4

hydrocarbon lean shale gas.2 To this regard, ethanol is a very interesting platform molecule

due to its steadily increasing production from biomass.1 Two classes of catalysts have been

used for ethanol conversion to 1,3-BD, namely ZrO2-based and MgO/SiO2-based (Lebedev

catalyst).3 The former have thoroughly been investigated using a combination of

computational and spectroscopic methods4,5 while the latter lack suitable spectroscopic

characterization.3 The overall reaction mechanism on MgO/SiO2 is currently debated3,6–8

and several recent attempts have been made to elucidate it.6,9–12 These studies pointed

towards aldol condensation as the most energetically favorable C-C bond formation

mechanism, except for Chieregato et al. who suggested that C-C bond was formed via

interaction of ethanol/acetaldehyde through a stable carbanion intermediate.9 The latter

mechanism was suggested to take place on pure, basic MgO sites based on a combination

of infrared spectroscopy and theoretical DFT results.9 The rate-determining step was found

to be ethanol dehydrogenation6,11 since an efficient dehydrogenating site was not present

in MgO/SiO2 catalysts. This suggests that an effective catalyst must possess

multifunctional, i.e. acidic, basic and redox sites. MgO/SiO2 catalysts are promoted with

transition metal (oxides) to improve their dehydrogenation capability2,13–17 where the

choice of transition metal used as a promoter is determined by its dehydrogenation

capability.18–20 Au,21,22 Ag,23,24 and Cu25,26 have been utilized to enhance the 1,3-BD

yield.2,27,28 Zn is another promoter that has been utilized to improve the yield of 1,3-

BD.13,15,29–31 The promotional effect was reported to originate from the improved

availability of both Lewis acid sites and redox sites.3,15 While Au and Ag promoters present

Page 217: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

197

economic constraints due to their high costs, Cu and Zn are relatively inexpensive and

present an alternative for an efficient catalyst design. The work reported here provides

new insights on the structure and reactivity of these sites under operating conditions.

Several theoretical and ultra-high vacuum (UHV) studies have been conducted on

Cu-based catalysts to determine the structure of the active sites32–39 but very few under

operating conditions. UHV characterization and DFT revealed formation of isolated or

clustered Cu0 phases on the MgO surface32,33 or a solid solution that contains Cu-Mg and

Cu-O-Mg bonds.34 For instance, on a perfect MgO (100) surface, DFT calculations showed

that a single Cu adatom prefers to bond with a surface O atom with the possibility to

spillover Cu.32,33 Various cluster sizes of Cu (dimers, trimers, and tetramers) were observed

depending on the surface coverage.32 The formation of reduced Cu clusters on the surface

was confirmed by Colonna et al. where Cu clusters, as evident by Cu-Cu bond length (2.55

Å), were observed as a thin layer on MgO using XANES during the UHV evaporation-

deposition synthesis.35 In a separate study, in addition to the observed Cu atoms on the

MgO surface, both UHV XANES and DFT identified the formation of a solid solution

between Cu and MgO that decreased the reactivity of the catalyst toward H2S and SO2

decomposition when compared to the supported Cu atom.36,37 Larger charge transfer

resulting in a strong ionic bond was observed when Cu was coordinated next to a defective

MgO surface.38,39 This shorter bond was due to the electron stabilization provided by the

Cu atom.38,39 UHV XANES of several transition metal-promoted MgO catalysts utilized

for CH3OH and RCH2Z (where R=H and CH3, Z=CN, COR’, and COOR”) coupling

reactions confirmed the formation of Cu-MgO solid solution at 80 K and suggested that an

octahedral coordination of the Cu species due to the pre-edge peak associated with 1s3d

Page 218: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

198

transition was very small. This observation was accompanied by the extended X-ray

absorption fine structure (EXAFS) analysis of the Cu-O and Cu-Mg atomic distances, 2.01

Å and 2.98 Å respectively, suggesting the formation of solid solution between Cu and MgO.

Interestingly, all promoted MgO catalysts that showed worse catalytic activity toward the

coupling reaction of CH3OH and RCH2Z (R=H and CH3, Z=CN, COR’, and COOR”) were

those that formed a solid solution with MgO.34 Applicable to this study is in-situ (operando)

characterization of a Cu-promoted catalyst for relevant alcohol reactions, such as methanol

formation from syngas,40 ethyl acetate production from ethanol41 and ethanol steam

reforming.42 Cu-containing ternary oxide catalysts, e.g. Cu/ZrO-SiO2, CuMgAlOx and

Cu/MgO-SiO2, were utilized for these and were well characterized.28,43,44 Operando and

in-situ characterization of these supported catalysts showed that Cu species could be

present as both Cu2+ ions and CuO - the latter exhibiting lower-strength interaction with

the SiO2 support,25,26 as a solid solution in the case of Cu-MgO/SiO228 and

Cu/ZnO/Al2O3,45 as dimeric structures in the case of CuMgAl hydrotalcite catalysts44 or as

reduced species as in the case for CuCrOx and CuZrSiOx catalysts.43,46 This suggests variety

of active copper sites can be present under operating conditions28,43–45 but very few studies,

notably Angelici et al.,26,28 attempted to decouple their reactivity during 1,3-BD formation

or investigate the temperature effect on Cu site composition under reactive conditions.28

ZnO/SiO2 has been used as a model catalyst for many reactions, such as water-gas shift and

methanol formation reaction,47 but X-ray based catalytic site characterization during

ethanol-to-1,3-BD are not existent to the best of our knowledge.13,15,16 In-situ XAS and

UV-Vis of this catalyst further showed the relevance of the precursor drying steps during

the synthesis and that Zn was present both as a silicate (hemimorpite) and ZnO bulk phase

Page 219: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

199

at 10% Zn loading.47 Ambient UV-Vis and TEM studies of a 1% ZnO/MgO catalyst

demonstrated the formation of a highly-dispersed ZnO layer which had high activity for

CO oxidation, affected by the quantum-confinement effect.48

In this work, we performed a comprehensive characterization on both Cu- and Zn-

promoted MgO/SiO2 catalysts. Details on the acid-base sites upon promotion with Cu and

Zn, implications for the reaction mechanisms, as well as thorough infrared, UV-vis,

electron and X-ray-based analysis of Cu and Zn local structure before, after, and during the

reaction was elucidated. Complementary, if not contradictory, conclusions were reached

for Cu-promoted MgO/SiO2 to those available in the literature28 while completely new X-

ray data insights were obtained for Zn-promoted MgO/SiO2 catalysts under operating

conditions.

2. Computational results

2. 1. Model catalyst selection and analysis

Plenty literature of Cu-doped MgO catalysts characterization is available, both on

computational and experimental studies. DFT calculation of small Cu cluster supported on

a perfect MgO (100) surface revealed that for a single Cu adatom, the preferential

adsorption site was on top of an O atom, whereas adsorption on a hollow site represented

a saddle point for Cu spillover.32,33 For a dimer Cu, there were two minimum states

available with close optimized energies, parallel and linearly perpendicular to the surface.

Two states were observed for the first case; one configuration where the dimer bond is 2.25

Å (stretched from 2.25 Å in its free form), and another one where the bond length is

stretched even further, 2.34 Å. The linearly perpendicular states, however, had the single

Page 220: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

200

adatoms’s Cu-O bond length, and free dimer’s Cu-Cu bond length. Trimer Cu and tetramer

Cu clusters preferred linear and rhombus geometry, respectively.32 Cu/MgO DFT model

had been used by Jose Rodriguez and coworkers.36,37 When Cu was embedded into the

surface, substituting an Mg atom with lower coordination, the catalyst was less reactive

compared to when the Cu atom was adsorbed freely on the surface. When a SH or S

molecule was adsorbed on the embedded Cu atom, the interaction was so strong that it

pulled the Cu atom out of the surface plane.36,37

A significantly larger charge transfer was observed by Zhukovskii et al. and

Matveev et al. when the metal was adsorbed on a defective MgO surface, meaning that the

bonding was more ionic than that on the perfect surface.38,39 The distance between Cu atom

and the surface had decreased as well, from ~2 Å to 1.62 Å. This stronger bonding

originated to the lower coordination atom, which would behave more like ions due to the

lack of electronic relaxation. The defect sites used here are both Fs and Fs+ sites, where an

oxygen atom was removed from a perfect surface, along with a number of electrons

accordingly to create the oxygen vacancy.

Experimental data carried out by Asakura and Iwasawa provided a different

insight.34 On a doped MgO catalyst, prepared using wet impregnation method, XANES

spectra suggested an octahedral coordination for Cu+ ions, deduced from the fact that the

pre-edge peak of the spectra which was assigned to the 1s-3d transition was very small.

The EXAFS spectra for Cu+ ion, revealed that the M-O and M-Mg distances were observed

to be 0.201 nm and 0.298 nm, respectively, away from the lattice constant of MgO, which

further suggested that the Cu+ ion would substitute an Mg site, i.e. supplanted into the

lattice.34 Colonna et al.35, however, failed to replicate the XAS experiments when the

Page 221: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

201

surface was prepared using ion evaporation-deposition method in an UHV chamber. On

monolayer coverage, the EXAFS data evidenced a Cu-Cu distance close to that of the bulk

metal, due to the weak film-substrate interaction. Copper was also observed to be in its

reduced state, and further, XANES spectrum showed a coordination number similar to the

bulk value, indicating that the Cu+ ions grew as a cluster on the MgO substrate.35 Different

preparation led to different geometry, as observed by Pascual et al., which carried out X-

ray measurement on Cu-doped MgO using arc fusion method, where MgO was melted

before the dopant is mixed as CuO.49 Both EXAFS and Ab initio calculation showed that

the crystal is in D4h geometry, associated with a compression of the original octahedron

along the z-axis, indicating that the ions substituted Mg sites in the lattice.

Although the literature on Cu/MgO catalysts is very well-established, very limited

study is available on promoted MgO/SiO2 catalyst. Complications on how Cu would be

added to the catalyst were brought upon by the presence of a second support material, i.e.

SiO228 or Al2O3

50. The Cu could be present either as a surface species on SiO2, as in the

case of Cu/SiO2 catalysts28, or as a substitutional dopant, replacing Mg as in the case of the

Cu/MgO catalysts. MgO/SiO2, the Lebedev catalyst, was studied by Angelici et al.28 Ex-

situ XANES, EXAFS, FTIR, XRD, TEM, XPS, and UV-Vis showed that Cu species were

not in planar geometry and were suggested to be located at crystal lattice sites. Formation

of small (CuO)x clusters on MgO-containing materials was also advocated experimentally.

XANES and EXAFS had also demonstrated the octahedral coordination of CuO and Cu-

Mg bond distance of 0.298-0.302 nm, similar to the Mg-Mg distance in periclase phase.28

Hydrotalcites (MgxAlyOz) were another class of catalysts that were routinely studied for

ethanol upgrading. High resolution NMR study of Cu-promoted MgxAlyOz was extensively

Page 222: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

202

studied and revealed that on low loading, Cu preferentially substituted for lattice Mg in the

hydrotalcite structure, while at higher Cu content, the transition metal was also present on

the surface as a bulk oxide.50

Studiy on the routinely used ZnO/SiO2 catalysts showed that zinc was mostly

present as Zn silicates in addition to small amount of ZnO bulk phase on the surface.47

Extensive characterization of the catalyst was carried out with XAS, DRIFTS, and UV-Vis.

Depending on the drying temperature, the small amount of bulk ZnO phase was formed on

top of the catalyst after calcination. The calcination mostly resulted in the formation of Zn-

silicate, hemimorphite in particular.47 However, when MgO was added to the catalyst, i.e.

Zn-doped Lebedev catalyst, more possibilities are now available, including the

substitutional doping of Mg by Zn, formation of ZnO bulk phase on either MgO or SiO2,

surface species formation on either support material, or preferential formation of Zn-

silicate. Colloidal suspension synthesis method of 1% ZnO/MgO was shown to result in a

very highly dispersed ZnO layer on top of MgO support, as confirmed by UV-Vis and

TEM.48 However, the catalyst synthesized in this study did not show the characteristic ZnO

band gap and operando XANES-EXAFS characterization of the catalyst suggested that the

local structure environment is very similar to Cu, indicating interaction with MgO, instead

of SiO2, resulting in a solid solution (vide infra). Hence, the model selection for both

catalysts was chosen to be CuMgO and ZnMgO, with both transition metals to

substitutionally dope an Mg atom. These models serve as a first approximation to the

catalysts’ model, simplifying the SiO2 support effects, further eliminating the contribution

of Mg-O-Si linkages and the accompanying hydroxyl groups.

Page 223: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

203

Table 6.1 Different configurations tested for Zn(Cu)/MgO model catalysts. Various dopant

location was chosen between the top and second layer, and compared for energy and Bader

charge.

Scheme Configuration Dopant location

Cu-doped Zn-doped

Energy

(eV)

Bader

Charge

Energy

(eV)

Bader

Charge

Cu(Zn)-

1 Cu3C-top layer

-659.48

(0.00) 0.72

-658.64

(0.00) 1.05

Cu(Zn)-

2 Cu5C-top layer

-659.28

(0.20) 0.82

-658.42

(0.22) 1.05

Cu(Zn)-

3 Cu4C-top layer

-659.22

(0.27) 0.82

-658.44

(0.20) 1.09

Cu(Zn)-

4

Cu5C-second

layer

-658.97

(0.51) 0.94

-658.15

(0.49) 1.16

Cu(Zn)-

5

Cu5C-second

layer

-658.89

(0.60) 0.92

-658.10

(0.54) 1.08

Cu(Zn)-

6 Cu4C-top layer

-659.13

(0.36) 0.81

-658.37

(0.27) 1.05

Cu(Zn)-

7 Cu4C-top layer

-659.45

(0.04) 0.91

-658.52

(0.12) 1.08

Cu(Zn)-

8

Cu5C-second

layer

-658.83

(0.65) 0.92

-658.09

(0.55) 1.08

Page 224: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

204

Cu(Zn)-

9

Cu5C-second

layer

-658.73

(0.75) 0.96

-658.03

(0.61) 1.13

Cu(Zn)-

10

Cu5C-second

layer

-658.75

(0.73) 0.93

-658.08

(0.57) 1.15

Table 6.1 shows the permutations tried for both dopants, i.e. Cu and Zn. The

original kink model from Chapter 3 was used and modified with dopants substitutionally

dope the catalyst’s surface at different Mg atom locations. From all of the tried models, Cu

(Zn)-1 possesses the lowest electronic energy. The Bader charge for the transition metal

atom for each configuration was also calculated, with charge values of +0.72 and +1.05 for

Cu and Zn. To discuss the effects of both Cu and Zn on the MgO model catalysts, Bader

charges of the neighboring atoms were calculated as well and compared to the undoped

MgO catalyst, shown in Figure 6.1.

Figure 6.1. Local structure analysis of (a)MgO, (b)Cu-MgO, and (c)Zn-MgO. The

Bader atomic charge on each atom is indicated by the boldfaced numbers.

(a) (b) (c)

Page 225: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

205

The oxygen atoms neighboring both transition metals exhibited less negative

charges than that of unpromoted MgO, with total charges of the 3 O atoms amounting to -

4.91, -4.31, and -4.4 e for MgO, Cu-MgO, and Zn-MgO, respectively. The decreased

atomic charges of O atoms neighboring the corner atoms indicates that the introduction of

both transition metal atoms had lowered the basicity of the O atoms. The lowered basicity

of the oxygen atoms neighboring the transition metal atoms was also observed both

computationally and experimentally using XPS on Zn-promoted talc by Baba’s group.13

The neighboring Mg atoms and other oxygen atoms that are distanced from the transition

metal dopants did not show any difference from the pristine MgO, which suggests that this

lowered basicity effect is very localized, and hence calculation should be focused on this

region.

The atomic charges observed for both Cu and Zn atoms are similar to what was

previously observed on CuO51 and ZnO52 surfaces. Bader charges for CuO and ZnO are

typically +0.57 – +0.84 and +1.13 – +1.20, respectively.51,52 The very positive atomic

charges for both means that these transition metal atoms are almost fully ionized in these

model catalysts. The lowered atomic charges of both Cu and Zn from the ionized Cu+1 and

Zn+2 are due to the charge transfer from the neighboring O atom. For Zn, the charge transfer

is very straightforward, since both Mg and Zn possess +2e charge. The charge transfer for

the corner Mg (Zn) from clean MgO (Zn-MgO) surface is equal to the difference between

the Bader charge value and the charge of isolated Mg (Zn). These values are 0.4e and 0.95e

for Mg and Zn, respectively. The difference of this value is equal to the difference of the

summed charges on the neighboring O atoms between MgO and Zn-MgO, i.e. -4.91 and -

4.4. The origin of charge density change is also acknowledged by Baba’s group for the case

Page 226: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

206

of Zn/talc.13 For Cu, however, this analysis fell apart, since the atomic charge of Cu is +1e,

while Mg is +2e, and the lowered total charges on neighboring O atoms is different from

the difference between Mg and Cu atomic charges.

2. 2. Reactive intermediates

Zn-1A Zn-1B (TS) Zn-2A Zn-2B (TS)

Zn-1C Zn-2C

Zn-3A Zn-3B (TS) Zn-4A Zn-4B (TS)

Zn-3C Zn-4C

Page 227: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

207

Cu-1A Cu-1B (TS) Cu-2A Cu-2B (TS)

Cu-1C Cu-2C Cu-2D

Cu-3A Cu-3B (TS) Cu-4A Cu-4B (TS)

Cu-3C Cu-4C

All optimized structures, including maxima and minima, are shown in Figure 6.2,

while the corresponding electronic energies and Gibbs corrected free energies were

tabulated in Table 6.2. The structures were optimized on the model Zn and Cu-doped MgO

Figure 6.2. All stable intermediates and transition states calculated following the reaction

pathways. (1A-1C): ethanol dehydrogenation to acetaldehyde; (2A-2C): ethanol

dehydration to ethylene; (3A-3C): C-C bond formation step in acetaldehyde aldol

condensation to 3-hydroxybutanal (acetaldol); (4A-4C): C-C bond formation step in Prins

condensation of acetaldehyde and ethylene. Calculations are carried over Zn/MgO model

catalysts (prefix: Zn), and Cu/MgO model catalysts (prefix: Cu).

Page 228: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

208

catalysts. To produce a comparable result, the optimization was done on the same site, i.e.

corner Zn (Cu) atoms and the neighboring O atoms. The reaction steps that were studied

are the key steps in both Prins and aldol mechanisms, i.e. dehydrogenation, dehydration,

and C-C bond formation of both aldol and Prins condensation pathways.

Table 6.2. Referenced electronic and corrected Gibbs free energy for each species over

MgO, Cu/MgO, and Zn/MgO catalysts.

For the dehydrogenation step, the intermediates are shown by Cu(Zn)-1A to 1C.

On all MgO, Cu-MgO, and Zn-MgO, ethanol undergoes spontaneous dissociation to give

surface ethoxide and a hydroxyl group. There was no visible difference from the model,

and the electronic energies are not significantly different, i.e. ~35-40 kcal/mol. The Gibbs

corrected free energies, however, are very different with Cu-1a is now the most stable

species and Zn-1a being the least. The transition state for the dehydrogenation reaction did

look very similar to the optimized structure on bare MgO catalyst (Chapter 3); the H-H and

H-lattice O4C distances are all of similar values.6 The energetic values, however, differ a

Species MgO Cu/MgO Zn/MgO

E

(kcal/mol)

G

(kcal/mol)

E

(kcal/mol)

G

(kcal/mol)

E

(kcal/mol)

G

(kcal/mol)

1a -40.6 -13.50 -37.33 -18.72 -35.75 -8.10

1b 3.9 26.20 14.33 32.12 12.36 36.23

1c -16.6 5.60 6.03 14.22 4.93 18.59

2a -39.9 -10.50 -25.66 -1.96 -31.45 -4.30

2b -2.6 23.00 4.97 21.48 5.99 31.54

2c -32 -15.30 -26.19 -1.65 -26.45 -8.22

3a -22.5 -23.70 -5.40 -14.32 -4.41 -4.21

3b -15.9 -7.60 3.07 3.22 -4.24 0.75

3c -19.4 -10.80 -2.89 -2.45 -8.72 -3.51

4a 11.6 -0.20 21.13 7.78 21.65 13.80

4b 25.6 28.60 34.17 31.13 33.81 35.49

4c -34.8 -22.90 -0.31 9.29 -27.25 -15.61

Page 229: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

209

lot from the bare MgO, ~6-10 kcal/mol higher than the transition state found for the bare

MgO. The very high values for both electronic and Gibbs’ corrected energies indicate that

these structures are very unstable. This observation is followed as well with the final state

of the reaction, i.e. the adsorbed acetaldehyde and hydrogen as a product of the reaction

when hydrogen is fully formed.

Ethylene production is an unwanted side reaction that accompanies 1,3-BD reaction

pathway, and was even linked to as a reactive intermediate in the Prins mechanism.53 Over

MgO catalyst, ethanol to dehydration possessed lower activation energy than

dehydrogenation, which confirms the experimental evidence.6 Over Zn(Cu)-MgO, the

initial state of this reaction, i.e. ethanol chemisorption, is very similar to the state observed

on MgO catalyst. Common structural parameters, such as the supposedly elongated C-O

bond, Cu(Zn) distance from the ethoxide O atom, and the bonding between O and planar

Mg5C are very similar to those in the bare MgO catalyst. Energetically, these intermediates

are much less stable than the MgO-bound species, differing ~8-10 kcal/mol. These

instabilities were also observed in the transition state and the final state of the reaction. Cu-

MgO, in particular, exhibits a very high degree of affinity toward the ethylene product, as

shown by the stabilized carbanion in Cu-2C. This intermediate then undergoes a C-Cu bond

breaking transition step (not optimized here) to give off ethylene as the final product, i.e.

Cu-2D. Cu-2D possessed referenced electronic energy of -27.54 kcal/mol, only 1 kcal/mol

more stable than Cu-2C.

Probably the most debatable step in the reaction mechanism is the C-C bond

formation. Aldol condensation had been widely accepted as the most possible mechanism,

and we optimized all the transition states over the doped model catalysts, following the

Page 230: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

210

calculated structure over bare MgO.6 Over the doped catalysts, the intermediates, i.e.

minima, are all less stable than that on MgO. The stability achieved by the system when

two acetaldehyde molecules are coadsorbed, i.e. initial state, on the surface is not replicated

well on the doped MgO. Surprisingly, the activation energy of this step is much lower for

Zn-MgO than Cu-MgO and bare MgO (discussed in detailed manner in Section 2.3). The

viability of Prins mechanism again is questioned computationally. On all the optimized

structures, similar activation energies were observed, i.e. ~21-29 kcal/mol. The

computational method also favors the reaction step thermodynamically, with all catalysts

giving exergonic reaction for the Prins step. This Prins mechanism, however, should be

treated carefully, since experimental evidence pointed out that ethylene formation is very

exclusive from 1,3-BD formation (Chapter 5).

2. 3. Potential energy surfaces

The potential energy surfaces for all computed reaction steps are presented in

Figure 6.3 and 6.4, while the activation barriers and Gibbs’ free energies of reactions are

tabulated in Table 6.3. Based on our calculation, promotion with transition metals, i.e. Cu

and Zn, did not achieve the intended lowered activation energy of ethanol dehydrogenation.

Rather, these activation barrier increased for the case of Zn and Cu to 44.33 and 50.84

kcal/mol, respectively. For dehydration, dehydration activation energies are consistent with

bare MgO, with similar values attained, especially for Zn-MgO. With the presented

activation energies for both C2 reactions, i.e. dehydrogenation and dehydration, the

reaction mixture at low temperature, will mostly consist of ethylene, instead of

acetaldehyde, for all catalysts.

Page 231: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

211

Table 6.3 Activation energy and thermodynamics consideration for key steps during

ethanol conversion to 1,3-butadiene over MgO, Zn/MgO, and Cu/MgO catalysts.

Reaction steps ΔGA (kcal/mol) ΔGRx (kcal/mol)

MgO Zn/MgO Cu/MgO MgO Zn/MgO Cu/MgO

Dehydrogenation 39.60 44.33 50.84 24.61 26.68 32.94

Dehydration 33.47 35.84 23.44 -4.87 -3.92 0.31

Aldol C-C 16.10 4.96 17.55 19.07 0.70 11.87

Prins C-C 28.75 21.69 26.39 -22.74 -29.42 -1.87

Very surprisingly, aldol condensation is much more favorable over Zn-MgO, with

the very low activation energy, as well as the lowered Gibbs’ free energy of reaction value

of 0.70 kcal/mol, which is almost aergonic. This lowered energetic barrier might be due to

the duality of Zn, which presents as both redox site and as a Lewis acid site, which

supposedly boost both dehydrogenation and aldol condensation step during the reaction.3

Another unexpected observation is that of Prins mechanism, which continues to show

viability computationally. Although this step does not have experimental ground, the

theoretical calculation shows that this step is very feasible. According to our calculation,

this step and ethanol dehydration are the two steps that have net positive rate constant

(Chapter 3).6 However, the calculation was carried out over model MgO catalyst, without

considering the presence of SiO2 and OH groups, therefore eliminating the other possible

sites, such as the open and closed Lewis acid sites of Mg-O(H)-Mg and Mg-O(H)-Si, which

were shown to actively catalyze the whole reaction steps (Chapter 5).54

The non-existence of hydroxyl groups in this idealistic model leads to very strong

Lewis acid-base pairs, which is accentuated by the electron-deficient sites such as kink,

corners, and edges. These highly unstable sites are typically stabilized by hydroxyl groups,

and leads to softer acid-base pairs, due to the more distributed electron density.

Page 232: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

212

Experimentally, the ethanol dehydrogenation step requires a weak basic site on the catalyst,

as shown by the CO2 and propionic acid cofeeding experiments, in-situ titration with CO2

and pyridine, which is also supported by previous investigation.2 The strong acid-base pairs

used throughout the calculation inevitably stabilize electron rich or deficient structures,

exhibited by the stabilized ethylene-acetaldehyde transition state that leads to the formation

of a C4 oxygenate.6 Furthermore, the proton abstraction steps that follow resulted in a

carbanion, which is stabilized by the presence of corner Mg3C2+, which is undoubtedly a

very strong electron acceptor.6

Another concerning discrepancy with experimental results are the dehydrogenation

step. The reaction mechanism is widely believed to be dictated by ethanol dehydrogenation.

The subsequent aldol condensation, dehydration, and MPV reduction steps were shown to

be facile and spontaneous on unpromoted MgO/SiO2 catalysts.11 Promotion with Cu14 and

Zn,15 among other transition metals,2,16,27 are intended to lower the ethanol

dehydrogenation step and to shift the rate-limiting step to MPV reduction step, which is

already considerably fast. The calculated activation energies for this step on the Cu-MgO

Figure 6.3. Potential energy surface for ethanol (a)dehydrogenation and

(b)dehydration over MgO, Zn/MgO, and Cu/MgO catalysts. (●)MgO, (■) Cu-MgO,

(♦) Zn-MgO.

Page 233: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

213

and Zn-MgO are, however, very similar to the bare MgO catalyst. On Cu-MgO, the

activation energy is even ~10 kcal/mol more than the bare MgO. There are two possible

sources of disagreement that is inherent to the model selection and the calculation nature.

DFT calculations on transition metals, especially first row transition metals such as Cu and

Zn, have to be treated carefully. Different DFT functionals had led to large variations in

energies, of 20 kcal/mol or more.55 The pure DFT functional used in this calculation (PBE)

is known to overestimate the stability of low-spin forms. Improvement is usually achieved

by including HF exact exchange, in expense of the prohibitively expensive calculation time,

especially in the large system used for this study.55 The assumption that the whole reaction

steps are carried out on one site is oversimplification of this complex system. While this

might be true on bare MgO, dehydrogenation over transition metal-promoted typically

occurred on an isolated transition metal sites,2,13,28 and the subsequent steps are over the

Mg-O(H)-Si or Mg-O(H)-Mg, which will be shown in the next sections.

Figure 6.4. Potential energy surface for first C-C bond formation via (a) acetaldehyde

aldol condensation and (b) Prins reaction between acetaldehyde and ethylene over

MgO, Zn/MgO, and Cu/MgO catalysts.

Page 234: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

214

3. Experimental results

3. 1 Catalyst characterization

The transition metal content in each catalyst was determined using both ICP-OES

and XPS to infer bulk and surface concentration, respectively. Interesting agreement was

found between the two characterization methods with ICP-OES determined Cu and Zn

content of 0.8 % and 2.5 % virtually agreeing with those determined by XPS of 0.9 % and

2.7 % for each catalyst. These Zn and Cu concentrations are close to the intended high

selectivity loading.14,15 The starting support material, i.e. wet-kneaded MgO/SiO2,

possessed surface area of 120 m2/g which was much lower than the fumed SiO2 used (332

m2/g). This lowering of the surface area has previously been observed by several other

groups14,27 and explained by the dispersion of low surface area MgO over SiO2. Promoting

the MgO/SiO2 samples with transition metals led to the increase in the surface area. Zn and

Cu-promoted samples exhibited surface area of 135 and 191 m2/g, respectively. This

significant enhancement of the catalyst surface area was not observed by Janssens et al.27

Rather, Ag-promoted samples were shown to considerably lower the surface area of their

calcined mesoporous support while in this study we used uncalcined hydroxide precursor.

This increase in surface area was likely due to the impregnation step which was done before

the support was calcined. The effect of calcination-impregnation order has previously been

observed by Da Ros et al. with ZrZn-promoted MgO/SiO2 catalysts.16 This suggests that

the metal promoters deposited via impregnation might act also as textural promoters, in

addition to being electronic promoters.

Page 235: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

215

X-ray diffraction (XRD) pattern of the two promoted catalysts – CuMgSi and

ZnMgSi – acquired under ambient conditions are shown in Figure 6.5 together with the

unpromoted MgSi. The unpromoted sample exhibited prominent peaks at 37.4, 43.5, 63,

75 and 79° which were due to the periclase MgO. Amorphous silica was also present in the

XRD pattern as evidenced by the broad band in the lower 2θ of 20-30° region. The wet-

kneading between MgO and SiO2 did not produce new bulk crystalline phases in agreement

with Angelici et al.54 Magnesium silicate hydrate phase was previously observed by

Shylesh et al. when MgO/SiO2 catalyst was synthesized by impregnating Mg precursor on

silica.2 Careful examination on the XRD pattern showed that Zn significantly enhanced the

intensity of the MgO peaks suggesting changes in its crystalline structure. For reference,

several concentrations of ZnSi and ZnMg were prepared and also analyzed with XRD

(Figure S6.1). ZnSi showed no new crystalline phases being formed up to 5% loading

while ZnMg also revealed no new crystalline phases were formed at loading up to 10%.

The Cu-promoted catalyst showed no change when compared to the support itself other

than the peak broadening of the MgO periclase structure. However, no new peaks appeared

Figure 6.5. Comparison of XRD patterns between CuMgSi, ZnMgSi, and MgSi.

Page 236: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

216

in the Cu-promoted sample as they would appear even at low loading on individual SiO2-

support (Figure S6.2).28 This also showed that Cu promoter was well dispersed on the

catalyst surface with no detectable oxide nanoparticle formed on the surface.28

Figure 6.6 shows DRIFT spectra for dehydrated metal-promoted catalysts in the

OH region, as well as that for the binary catalyst component compounds (ZnSi, ZnMg,

CuSi, CuMg). The promoted MgSi catalysts show similar spectral features to the

unpromoted MgSi. Detailed assignments of the four native OH groups can be found in the

previous work.11 Briefly, there are four prominent peaks on an MgO/SiO2 catalyst, i.e. 3745

cm-1 assigned to both isolated MgO and silanol groups, 3725 and 3705 cm-1 ascribed to

Mg-OH-Si with different OH coordination numbers and 3680 cm-1 peak assigned to a

magnesium silicate species. Promoting the MgSi with Cu or Zn significantly reduced and

broadened the native silica and the WK-signature peaks, i.e. isolated silanol at 3745 cm-1

and Mg-O(H)-Si group at 3680 cm-1. This suggests that both transition metal promoters,

Cu and Zn, interact strongly with this OH group as well. Displacement with Zn further

Figure 6.6. In-situ dehydrated DRIFTS of OH region of MgSi, CuMgSi, and ZnMgSi.

Spectra were taken at 100°C under He flow after pretreatment at 500°C for 1 hour.

Spectra were offset for clarity.

Page 237: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

217

results in a new OH site, as shown by the emergence of a peak at 3760 cm-1, which was

previously assigned to the isolated hydroxyl group of MgO.11,56 This highly isolated

hydroxyl group might form from broken Mg-O-Si linkages due to the introduction of Zn

suggesting Zn interaction with O-Mg.

The coordination and oxidation states of the metal promoters are further

characterized by in-situ UV-Vis DRS under dehydrated conditions. Figure 6.7a shows a

comparison between the Cu-promoted (CuMgSi) catalyst, MgSi and reference binary

materials, such as CuMg, CuSi and bulk CuO. UV-Vis DRS spectra of the bulk CuO is

characterized by the presence of a charge transfer (CT) peak at ~251 nm and a peak at 570

nm. The CT peak is assigned to the ligand-to-metal CT (LMCT) from O2- to Cu2+ in

octahedral coordination.44 The peak at 570 nm can be assigned to either surface plasmon

resonance from Cu0 or contributions from d-d transition.57 Furthermore, a peak at 235 nm

Figure 6.7 In-situ UV-Vis DRS spectra of (a) dehydrated CuMgSi catalyst referenced

with Cu/MgO (CuMg), Cu/SiO2 (CuSi), CuO, and MgSi; (b) dehydrated ZnMgSi

catalyst referenced with Zn/MgO (ZnMg), Zn/SiO2 (ZnSi), ZnO, and MgSi. Inset: UV-

Vis spectra of different loadings of Zn on MgO/SiO2 catalysts.

(a) (b)

Page 238: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

218

is present on all supported Cu samples, while the peak at 270 nm is present only on a Mg-

containing support. The former represents LMCT peaks for a very isolated Cu-O

species,28,44 while the latter has been assigned to an oligomeric Cu-O species.44

Analogously, the peak at 305 nm for CuSi is also assigned to the oligomeric Cu-O

species.28 This reference sample (CuSi, Figure 6.7a) also exhibits d-d transition peak at

~760 nm, indicative of Cu2+ species in a (distorted) octahedral field.28 On the other hand,

the CuMg reference exhibited an extra peak at 215 nm, possibly due to charge transfer

between Mg2+ to silica surface.27 This peak is not present on the CuMgSi catalyst which

means that Cu promotion eliminated this exposed Mg species, consistent with DRIFTS

observations. The CuMgSi catalyst exhibits a small peak at ~570 nm, which, as in the CuO

reference case, might be due to the presence of a reduced species, i.e. Cu2O or Cu0. Lower

geometry species are hardly encountered on mixed metal oxide and dehydration under inert

atmosphere is more likely to induce partial reduction on the catalyst.28 In agreement, a

known adsorption peak in the 560-570 nm region is due to the plasmon resonance of

metallic Cu nanoparticles.57

Tauc plots of CuO standard and the catalyst (CuMgSi) were derived from the UV-

Vis DRS spectra, shown in Figure S6.4. Using the method previously described by Bravo-

Suarez, et al.44, identification of the oligomer is made possible by correlating the number

of species to the edge energy. The plot in CuMgSi was deconvoluted into two species,

isolated (0 nearest neighbors) and the oligomer that will be determined, with edge energies

of 3.86 and 3.51 eV, respectively. The Tauc plot indicates that the reference oxide CuO

exhibits an edge energy of 1.26 eV, close to the previously determined values at 1.17 ±

0.06 eV.44,58 The value for the isolated CuO was higher than that reported for CuMgAl

Page 239: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

219

mixed oxide, which had been reported to be ~3 eV.44 The difference originates from the

coordination of the isolated CuO. In the previous report it was determined that the Cu

species is in the Cu2Al domains, instead of forming a solid solution with Mg.44 Using

isolated CuO species and standard CuO (6 nearest neighbors), the coordination number, i.e.

number of Cu-O-Cu bond, was determined to be 0.8.

The Zn-promoted catalyst UV-Vis DRS spectra are shown in comparison with the

reference samples, i.e., bulk ZnO, MgSi, ZnSi and ZnMg, shown in Figure 6.7b. The

ZnMgSi catalyst shows a small peak at 276 nm. This small peak is down shifted ~100 nm,

when compared to bulk ZnO at 360 nm. Additionally, ZnMgSi contains a peak at 215 nm,

which resembles that of the CuMg UV-Vis DRS spectrum. This CT peak appears in almost

all Mg containing samples, except for CuMgSi. That peak was located at almost the same

wavelength, ~215 nm, for CuMg, ZnMg, and ZnMgSi, but shifted when MgSi support was

used, i.e. at 225 nm. This peak can be assigned to a charge transfer from Mg2+ to O2-, where

a shift is expected when MgO is wet-kneaded with SiO2.59 However, introducing Zn to the

MgSi support seems to negate this shift and it reverts back to ~215 nm. This phenomenon

is consistent with DRIFTS data, as shown in Figure 6.5, where the OH peak at 3740 cm-1

disappeared when MgO was wet-kneaded to SiO2, but reappeared when Zn is introduced

to the surface. Figure 6.7b inset shows different Zn loadings on the wet-kneaded MgSi.

At a higher loading, the peak at lower wavenumber, i.e. 215 nm, persists, while the ZnO

peak started appearing at 270 and 280 nm for 10% and 15% Zn loadings, respectively. The

shift in the CT peak is also followed by the shift in the edge energy cutoff. This shift with

a higher Zn loading was also observed by Yoshida et al. on an SiO2 support, although they

Page 240: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

220

describe this Zn site to have a distinct electronic structure from bulk ZnO, with XANES

confirming that the ZnO is in a tetrahedral configuration.60

The reference ZnMg and ZnSi samples further aided in peak assignments of the

UV-Vis spectra of the ZnMgSi catalyst. In addition to the discussed 215 nm peak, the

former exhibits two other peaks at 276 and 360 nm. The first peak could be associated with

the defected Mg site of the catalyst, assigned to tri-coordinated O2- ions on corner sites,

which is also encountered in the MgSi sample.27,59,61 Along with the peak at lower

wavelengths, 215-225 nm, these peaks are indicative of the bulk MgO, also observed by

Figure 6.8. Scanning Transmission Electron Microscopy images of ZnMg, ZnMgSi,

CuMg and CuMgSi samples. Energy Dispersive Spectroscopy profiles (smoothed)

are also provided. Small ZnO nanoparticles are shown in ZnMgSi with red arrows.

Page 241: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

221

Sels and coworkers.27 The second peak is likely to be assigned to bulk ZnO as a compared

to the bulk ZnO reference spectra. The ZnMgSi catalyst, on the other hand, hardly shows

any other peaks related to Zn-containing species, which rules out the bulk ZnO from active

site consideration. Bulk ZnO nanoparticles might only be formed at a very small particle

size. This is supported by STEM measurements in Figure 6.8. Chouillet et al. reported a

similar observation, where UV-Vis shows bands of a bulk ZnO phase in the limit of 1.4-

4.4 nm particle size, confirmed by TEM.47 Highly dispersed ZnO nanoparticles have also

been previously observed on MgO-supported catalysts.48,62 Another possibility is that Zn

might be present in a solid solution inside the lattice of the support (vide infra), as

previously reported in SiO247,60 or in talc.13 In particular, ZnMg shows large ~30 nm

isolated ZnO crystals present. However, ZnMgSi shows very small ~1 nm crystals and the

presence of isolated ZnO nanoparticles. This is consistent with the UV-Vis data shown in

Figure 6.7. Isolated (monomeric) Cu sites, as well as oligomeric sites in both CuMg and

CuMgSi, can’t be detected using STEM/EDS in Figure 6.8, indicating high dispersion of

these sites.

To confirm the presence of some reduced species on the surface, oxidative

treatment was done post-inert treatment by flowing air (Figure S6.5). The significant

increase in the CT bands at 250 and 310 nm in expense of the peaks at 575 and 633 nm for

CuMgSi indicates the presence of some native reduced species that became oxidized upon

the introduction of air at higher temperature. Similarly, ZnMgSi shows the continuous

increase in peaks at 230 and 340 nm, indicating the formation of both MgSi sites and bulk

ZnO phases when oxidized.

Page 242: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

222

3. 2 Steady state catalytic performance and acid/base chemistry of the catalyst

active sites

The steady state reactivity comparison between MgSi, ZnMgSi and CuMgSi

catalysts is shown in Figure 6.9. Here the activity of three catalysts is compared in the

temperature range of 350-450°C. It can be seen that promotion with Cu and Zn significantly

enhanced the 1,3-BD formation rate from <1 mmol/gcat h to ~2 mmol/gcat h throughout the

investigated temperature range. Furthermore, ethylene formation was suppressed, more

significantly in the case of Zn promotion. The origin of this promotional effect can be

traced back to the production of acetaldehyde, which significantly increased in comparison

to the unpromoted catalyst. This accumulation of acetaldehyde on the surface indicates

that the Rate Determining Step (RDS) shifted for the case of promoted MgO/SiO2.

Quantitatively, this is confirmed by the decrease in apparent activation energy, Ea, as

derived from the Arrhenius plot of each product formation rates. Acetaldehyde and 1,3-

BD activation energy exhibits similar trend with promotion with Cu and Zn, with Ea (Zn)

< Ea (Cu) < Ea (unpromoted). Apparent activation energy of ethylene, on the other hand,

decreases with Cu promotion but not with Zn. With Zn, however, increasing temperature

does not increase the formation rate of ethylene, which explains very low activation energy

on this catalyst. The very low formation rate of ethylene must be due to very low rate

constant of ethylene formation, since raising the reaction temperature does not have

significant effect on the formation rate.

A similar increase in 1,3-BD production was previously reported by various

investigators.3 For instance, Weckhuysen and coworkers noticed a sharp increase (~20%)

in both ethanol conversion and 1,3-BD yield upon promoting the wet-kneaded catalyst with

Page 243: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

223

1% CuO. The productivity of their catalyst was very similar to that reported here: 0.48

mmol gcat-1 hr-1 at 425°C and WHSV = 1.1 hr-1.14 When the reaction was carried out at

more than 375 °C, the conversion over ZnMgSi approached 100%. This increase in

conversion was previously observed when Zn was shown to provide more Lewis acidity

and also suppressed the Brønsted acidity.15,63 Zn-promoted catalysts, such as MgO/SiO215

and talc13, were reported to increase both the conversion and selectivity toward 1,3-BD.

The latter showed the same productivity as our catalyst, ~1.1 mmol gcat-1 hr-1 at an even

lower reaction temperature (300°C) and a much higher WHSV (8.4 hr-1).

The change in the surface chemistry of the catalyst induced by the presence of these

metals, to the best of our knowledge, has not been thoroughly investigated. The general

consensus is that the catalyst should have all redox, basic, and acidic sites on its surface.

On Cu, extensive study on the local coordination of Cu by means of XAS was not

accompanied by the identification of molecular coordination by other spectroscopic

methods.28 Further, promotional effects on Zn-promoted MgO/SiO2 catalyst were not

extensively investigated, i.e. studies were only focused on the activity change and the

implication on acid-base characteristics of the catalysts.15 Basic site poisoning using CO2

and propionic acid can reveal the reactive site difference between the three catalysts. CO2,

a relatively weaker acid than propionic acid, will occupy the stronger basic sites2,64 while

propionic acid should non-discriminatively adsorb on all basic sites given its stronger

acidity. Coflowing CO2 with ethanol as a weak acid will mainly poison the strong basic

sites and suppress any reactions that require participation of these sites. Propionic acid,

being a stronger acid, will indifferently poison any basic sites, possibly suppressing all

detectable reaction products. When switching to reactant-only flow, the weak bond

Page 244: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

224

established between the weak CO2 molecules should be broken and therefore should

eliminate the poisoning effect and revert the system back to its original state. For propionic

acid, however, the strong basic sites should maintain strong bond with the probe molecule

after flow is stopped and should irreversibly deteriorate the productivity of one or more

reaction products that depend on the site availability.

Figure 6.9. Productivity comparison of 1,3-BD (■), ethylene (●), and acetaldehyde (▲)

over (a) MgSi, (b) CuMgSi, and (c) ZnMgSi. Dotted lines are meant to guide the eyes.

Insets: Arrhenius plots to show apparent activation energies of the three (by)products.

Reactions are carried out between 325 - 450°C, mcat = 0.1 g, pethanol = 1.8 kPa, total flow

= 55 ml/min.

Page 245: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

225

Fundamental acid-base study on both transition metal-promoted catalysts were

investigated by both in-situ and ex-situ methods (Section S6.2). In-situ studies using

propionic acid showed that all three catalysts possessed very limited amount of strong basic

sites and that promotion with transition metals further decreased the amount of strong basic

sites. The propionic acid cofeeding experiment showed that 1,3-BD productivity did not

recover to its original formation rate which suggests the presence of some strong basic sites

that maintain strong interaction with the leftover propionic acid.2 With the wet kneaded

support, the strong basic sites are limited and more medium basic sites are present. Both

in-situ CO2 poisoning and DRIFTS study confirmed the increased availability of the

medium and weak basic sites. Our study aligns well with previous study using deuterated

chloroform, with Cu-Mg solid solution being thought of as the origin of reduced strong

basic sites.28 The in-situ poisoning further unraveled the site requirements for every step of

the reaction, i.e. acetaldehyde formation on weak basic sites, dehydration on any sites, aldol

condensation and MPV reduction on strong basic sites. The reduced amount of strong basic

sites is also the origin of RDS shift from acetaldehyde formation to MPV reduction. Total

amount of acid sites were also reduced by promotion with Zn and Cu, as shown by both

in-situ NH3 poisoning and NH3-DRIFTS experiment. While acid sites are responsible for

the dehydration steps, the origin of acetaldehyde formation rate reduction is the competitive

bonding between the available Cu2+ to NH3, since Cu catalysts are routinely investigated

as SCR catalysts.65,66 This is further supported by the recovered acetaldehyde production.

The acetaldehyde production was accompanied by Cu2+ successive reduction to Cu0, as

shown by in-situ XANES (vide infra) and was possibly the reason its productivity

decreased overtime. Promotion with transition metals yielded similar results, where Lewis

Page 246: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

226

acid sites associated with Mg3C2+

is removed, with enhancements on the M4C2+ sites. This

finding indirectly confirms the structural change in the catalyst itself, i.e. solid solution

formation.

3. 3 Active sites under operating conditions

3. 3. 1. Temperature programmed infrared spectroscopy measurements (TP-

DRIFTS)

The effect of metal promoters on ethanol to 1,3-BD reaction mechanism were

probed using in-situ temperature programmed DRIFTS. This allowed to study the surface

species participating during the reaction. Detailed assignments of the IR peaks can be found

elsewhere.11 Briefly, experiments utilizing different probe molecules, i.e. ethanol,

acetaldehyde, crotonaldehyde, and crotyl alcohol, were performed. Table 6.4 summarized

the peak assignments from experiments done on MgSi catalyst. The in-situ DRIFT spectra

in 1700 to 1300 cm-1 region of MgSi, ZnMgSi and CuMgSi catalysts are shown in Figure

6.10 (insets). There were two very prominent peaks in the spectra at high reaction

temperatures (>250°C), i.e. ~1575 cm-1 and 1440 cm-1, previously assigned to the product

of acetaldehyde aldol condensation and polymerization.11 Noticeable difference between

the unpromoted spectra and the promoted ones was in the exact position of the two peaks.

On promoted catalysts, the C=C stretch shifted to 1587 cm-1 while the prominent peak for

the C-H bending was at 1458 cm-1. The 1587 cm-1 peak location is identical in the case for

both CuMgSi and ZnMgSi, which indicates similar anchoring site on the catalyst. As will

be discussed later some of the magnesium forms solid solution with both Cu and Zn, which

is possibly the binding site of the reaction product, given the identical peak location.

Page 247: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

227

The C-H bending peak was very complex since every reactive intermediate has a

C-H group. Peaks were deconvoluted using CasaXPS software suite version 2.3.18PR1.167

into several different components. On the unpromoted catalysts, this broad envelope was

deconvoluted into four peaks, i.e. 1458, 1440, 1416, and 1398 cm-1. On metal-promoted

catalysts, these peaks were less convoluted showing fewer species involved with only three

prominent peaks existing. Interestingly, the peak at 1458 cm-1 was formed more rapidly in

the case of promoted catalysts, while peaks at 1435 and 1416 cm-1 lagged, compared to the

unpromoted catalyst. The growth of the peak at 1458 cm-1, previously assigned to

acetaldehyde (δ CH3) and crotonaldehyde (ρw CH3), is significantly enhanced over

promoted catalysts. The reactive nature of acetaldehyde, which is the generally accepted

Figure 6.10 Evolution of each peak during in-situ temperature-programmed ethanol

DRIFTS over (a) MgSi, (b) CuMgSi, (c) ZnMgSi. Insets: original spectra of ethanol

DRIFTS from where the peaks were deconvoluted.

Page 248: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

228

first reactive intermediate, complicates analysis where multiple competing reactions, such

as aldol condensation, acetate formation, and polymerization to take place at low to

intermediate temperature.68–73 The adsorbed acetate formation can’t be fully ruled out due

to the peaks at 1587-1575 cm-1 that appeared and grew almost at the same rate with the

~1400 cm-1 peak.70 Together with polymerization of acetaldehyde and consecutive aldol

condensation to C6 aldehydes, these reactions present side reactions that occur. The acetate

formation is doubtful to take place in this experiment. In particular, if peak at 1587 (1575)

cm-1 is assigned to the surface acetate the change in the growth after promotion with Zn

(Cu) would apply to all the peaks in the 1460-1400 cm-1 region. In fact, the improvement

in growth of peak at 1458 cm-1 after promotion is much more significant than that for the

peak at 1587 (1575) cm-1.

Hence, the peaks at 1587-1575 cm-1 and 1457 cm-1 can be used to characterize the

degree of both aldol condensation and dehydrogenation that takes place on the surface,

while the other peaks at ~1400 cm-1 to characterize the catalysts’ basicity, i.e. its ability to

readily polymerize the formed acetaldehyde. The resulting crotonaldehyde tends to stay

on the surface and further undergo other reaction than to desorb as vapor-phase

crotonaldehyde. The C4 intermediate can be further aldolized with acetaldehyde to form

2,4-hexadienal and stick on the surface and possibly deactivate the catalyst.74 This insight

can be further utilized to probe the abundance of the active sites of the catalyst, i.e. based

on the accumulated 2,4-hexadienal which was characterized by the 1587 cm-1 peak. We

carried out semi-quantificative analysis of the peaks at 1587 (1575), 1440, and 1458 cm-1.

The peaks at ~1400 cm-1 are summed together assuming that they result from similar class

of reaction, i.e. polymerization that typically yield more than one product such as

Page 249: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

229

metaldehyde and paraldehyde.74 The evolution of these peaks as a function of temperature

was plotted in Figure 6.10. It can be seen that for all catalysts, there was no significant

changes in the ~1400 cm-1 peak area. However, the promoted catalysts resulted in a higher

intensity/area of the 1587 cm-1 peak with Cu higher than Zn. This indicates that promoting

the catalyst with transition metal promoters enhances the ability of the catalyst to carry out

aldol condensation, while at the same time keeping the unwanted polymerization constant

with regards to the unpromoted catalyst. Another noticeable difference was the temperature

where the peak started increasing in intensity. For Cu, the peak starts increasing at lower

temperature, even at ~150 °C, while Zn lagged behind and showed similar reactivity to the

unpromoted catalyst.

Overall, combination of both DRIFTS and steady state fixed-bed experiments

showed a shift in the rate-limiting step. Without the promotion with transition metal, less

acetaldehyde was produced in the product stream indicating the rapid consumption of the

intermediate. Promoted catalysts, on the other hand, saw increase in acetaldehyde

production, which suggested a bottleneck reaction. The accumulation of acetaldehyde in

the steady-state reaction experiments suggested that aldol condensation is the RDS. The

acidity and basicity of the catalyst was affected as well by promotion with transition metal.

In-situ poisoning experiment with propionic acid and NH3 showed that promoting increases

the availability of the weak basic sites and total acid sites, as shown by the significant

decrease in the production of all products during the coflow. In-situ DRIFTS of ethanol

over the three investigated catalysts indicated that there was a change in the binding site

during the aldol condensation, as manifested by the shift of C=C stretch peak at 1575 to

1587 cm-1. This systematic change suggested that while the anchoring site was identical

Page 250: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

230

between the two promoted catalysts, a potential solid solution formation took place.

Mechanistically, this semi-quantification confirms the steady-state experiment findings

where the activation energy of the dehydrogenation step was significantly reduced leading

to higher amount of acetaldehyde and products of aldol condensation. The change in the

polymerization products was also an indication to the altered basicity of the catalyst.68,69

Though the difference was not significant, the reduced polymerization products indicated

that the basicity of the catalyst was slightly reduced.

Table 6.4. Vibrational frequencies in 1600-1400 cm-1 wavenumber range and their

assignments for ethanol, acetaldehyde, crotonaldehyde and crotyl alcohol adsorption on

WK (1:1)11

Assignment

Experimental (cm-1)

Ethanol Acetaldehyde Enolate Crotonaldehyde Crotyl

alcohol

ν (C=C) - - 1600,

1578 1600, 1574 1602

δ (CH2) 1454 - - - 1380

δ (CH3) 1418 - - 1456, 1434 1368

ρw (CH) 1380 - - - -

ρw (CH2) - - - - 1441

ρw (CH3) 1338 1456, 1434,

1382 - 1346 1456

3. 3. 2. In-situ UV-Vis DRS study of MgSi catalysts

Figure 6.11 shows the in-situ UV-Vis DR spectra during ethanol conversion to 1,3-

BD on (a) CuMgSi and (b) ZnMgSi. The spectra plotted are difference spectra referenced

to 100 °C to better describe the dynamic changes. On CuMgSi it can be seen that with the

reaction progressing there were four broad spectral bands. Increasing the temperature lead

to the intensity increase at 248, 315 and 565 nm while the band at 276 nm showed decrease

in intensity. Interestingly, inset in Figure 6.11a shows that the band at 211 nm reached a

Page 251: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

231

maximum at 300°C and decreased in intensity at higher temperature. To assist the peak

assignments, we performed similar experiments on unpromoted MgO/SiO2 catalyst

(Figure S6.12). The UV-Vis spectra of the unpromoted catalysts showed changes on three

bands at 210, 245, and 300 nm. These three peaks can be assigned to either CT bands of

metal oxides, π- π* transitions of allylic cations, cyclic or aromatic species, or even neutral,

uncharged aromatic species (for shorter wavelengths).75,76 An alternative assignment for

the two bands at 210 and 245 nm was the LMCT band of Mg to O on defect sites and to

SiO2, respectively.27,59 The remaining peaks are at 276 nm that decreased at the expense

of peak at 565 nm. The former was assigned to oligomeric CuO species (~0.8 Cu nearest

neighbor), while the latter one was assignable to surface plasmon resonance of Cu also due

to the rare occurrence of lower geometry CuOx species in mixed oxide systems.28,44 The

indicated reduced CuO oligomeric species to surface Cu0 will later be confirmed by X-ray

methods since peak at 565 nm could also originate from substituted or unsubstituted

benzene.75

On ZnMgSi, in-situ UV-Vis experiments showed the emergence of different

intermediates as signified the by the bands at ~240 -shifted to 268 nm at higher

(a) (b)

Figure 6.11. In-situ UV-Vis DRS under constant ethanol flow over (a) CuMgSi and (b)

ZnMgSi

Page 252: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

232

temperature-, 300 and 211 nm. These bands were observed on the in-situ ethanol

experiment over unpromoted MgO/SiO2 catalyst as well. Another band with a cutoff at 350

nm also appeared. This band could be indicative of π- π* transitions of dienic allylic

cations75 or bulk ZnO formation since its emergence was also accompanied by the intensity

increase of shoulder at ~230 nm, which alternatively can be assigned to CT between Mg2+

to SiO2.27 The alternative assignment can suggest that Zn was transformed from its solid

solution state into bulk ZnO, as accompanied by the formation of CT band at ~230 nm.

3. 3. 3. Operando XAS studies of Cu, Zn-promoted MgSi catalysts

3.3.3.1. Operando XANES and EXAFS of Cu-promoted MgSi catalyst

The XANES spectra of Cu catalysts and standards taken under ambient condition

are shown in Figure 6.12. The XANES spectra for samples with Cu-promoted supports,

i.e. CuMg, CuSi, and CuMgSi, show similar features with a weak pre-edge peak located at

about 8977 eV and a shoulder peak at the rising edge at about 8987 eV (Figure 6.12, left).

The weak feature at 8977 eV was previously assigned to the 1s → 3d transition, and is

Figure 6.12. Normalized XANES spectra of CuMg, CuSi, and CuMgSi (a) and Cu foil,

CuO, Cu2O, and CuMg (b). Inset: Cu K-edge k2-weighted EXAFS data of

corresponding spectra. XANES spectra in Figure 6.12(a) were offset vertically for

clarity.

Page 253: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

233

considered a signature for Cu2+ species.28,77,78 For comparison, XANES spectra of the

standards, i.e. Cu foil, Cu2O, and CuO, are plotted along with CuMg XANES spectrum.

The CuMgSi catalyst XANES spectrum strongly resembles that of the CuMg, and is very

different from CuSi and Cu standards. Further, the EXAFS spectra in the inset are very

similar for both CuMg and CuMgSi. The shoulder peak at 8987 eV, when compared to

CuO, was shifted from 8985 eV. This shoulder peak is usually assigned to the 1s → 4p

transition, and its position is associated with neighboring atomic geometry.79 For CuMg,

the shift in the shoulder peak was also observed. 28 Many reports attributed that shift to Cu

being in octahedral or distorted octahedral geometry, occupying Mg lattice sites in a solid

solution.34,35,49

Table 6.5. Best fitting results of Cu catalysts. The structural parameters of standards were

listed for comparison.

Sample Bond N R (Å)

CuMgSi Cu-O 5.6±1.1 1.96±0.02

Cu-Mg 7.0±1.8 3.01±0.02

CuMg Cu-O 4.5±0.9 1.97±0.02

Cu-Mg 7.1±2.0 3.00±0.03

CuO

Cu-O 4 1.96

Cu-O 2 2.78

Cu-Cu 4 2.9

Cu-Cu 4 3.08

Cu-Cu 2 3.18

Cu2O Cu-O 2 1.84

Cu-Cu 12 3.01

MgO Mg-O 6 2.11

Mg-Mg 12 2.98

Cu foil Cu-Cu 12 2.56

As shown in Figure S6.13 (the Fourier transformed k2χ(k) spectra of CuMgSi,

Cu2O, CuO and Cu foil), the R-space EXAFS spectra of CuMgSi have two distinct peaks

Page 254: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

234

in the range of 1-3 Å. The peak at about 1.5 Å is due to Cu-O contribution, and the peak at

about 2.6 Å could be due to Cu-Cu contribution from Cu oxides or Cu-Mg contribution if

Cu enters MgO lattice. To determine the local environment of Cu, EXAFS analysis was

performed and two models were tested. Model A includes Cu-O and Cu-Cu path and Model

B includes Cu-O and Cu-Mg path. The fitting k range is 2.0-11.0 Å-1 and R range is 1.0-

3.1 Å. The best fitting results were obtained by using Model B and are shown in Table 6.5.

For comparison, the structural parameters for Cu foil, CuO, Cu2O, and MgO were also

listed in Table 6.5. The Cu-O bond parameters on both samples are similar to those of Cu-

O bond in the CuO. The Cu-Mg bond lengths in both CuMg and CuMgSi are also similar

to the Mg-Mg and Cu-Cu bond lengths of MgO and CuO standards, respectively. The Cu-

Cu contribution was not detected for either CuMg or CuMgSi, which corroborates the

insertion of Cu into MgO lattice. Coordination number of Cu-O shown in the EXAFS

analysis was also in line with the (distorted) octahedral geometry. Previous investigations

by Asakura et al. and Angelici et al. demonstrated that Cu-O coordination numbers were

lower than 6.28,34 Angelici, et al. found a coordination number of 4 and further assumed the

presence of two additional oxygen atoms to simulate the XANES spectra which revealed

another contribution from Cu-O bond at ~2.40 Å, which is characteristic of a separation

between copper and apical oxygen atom in a CuO6 complex.28 For CuMg, the Cu-O

contribution follows similar observation of Angelici et al. and Asakura et al., i.e. less than

6.28,34

Page 255: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

235

Operando XAS experiments with flowing ethanol over CuMgSi were performed at

different reaction temperatures to analyze the role of Cu species during the reaction, and at

400°C, multiple scans were performed to investigate the evolution of Cu species as the

reaction progresses at constant temperature. Figure 6.13 shows the XAS spectra of

CuMgSi under both helium flow (a) and constant ethanol flow (b) at different temperatures.

As shown in Figure 6.13, the pre-edge peak (at 8977 eV), which is a signature of Cu

divalent species, remains almost unchanged after pretreatment, indicating Cu remains in

the 2+ state after He treatment. Under helium at elevated temperatures, a new feature at

8982 eV appeared suggesting the change of the local environment of Cu after pretreatment.

The position (8982 eV) of this peak is quite close to that (8981 eV) of the shoulder peak of

Cu2O, in which each Cu atom is surrounded by two O atoms in a collinear manner. The

appearance of the 8982 eV peak thus implies the decrease of the average coordination

number of Cu-O bond for Cu atoms in CuMgSi catalyst. During experiment with ethanol,

significant increase in the intensity of 8982 eV peak was observed especially at high

Figure 6.13 Normalized temperature-programmed operando XANES spectra of

CuMgSi catalyst under He flow (a) and ethanol flow (b). Inset: enlarged region of the

pre-edge features to elucidate changes at different temperatures.

Page 256: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

236

temperatures, suggesting the increased fraction of species in which the average Cu-O

coordination number is low. We propose that such geometry is correlated with catalytic

activity of CuMgSi catalyst. The corresponding MS data (Figure S6.14), shows that the

acetaldehyde was made at very low temperature, i.e. starting as low as 100 °C, and

increased significantly at ~250 °C. This increase is reflected as well by the spectra at 300 °C,

where the increase is very significant from 200 °C. At the same time, the 1,3-BD started

being produced at ~250 °C, which was lower than unpromoted catalyst, i.e. 300 °C.

When reaction temperature reached 400 °C, the temperature was held constant

while XANES spectra were repeatedly taken to investigate any changes that take place

during the reaction. The change in the copper species was recorded as a function of time,

shown in Figure 6.14. A Cu foil XANES spectrum taken at ambient temperature was

overlaid for comparison. As reaction proceeded, the peak at 8982 eV started decreasing in

intensity, suggesting the re-arrangement of the local structure of Cu. Accompanied with

Figure 6.14. Normalized time-resolved operando XANES spectra of CuMgSi catalyst

under ethanol flow at 400°C. Inset: enlarged region of the pre-edge features to elucidate

changes at different temperatures.

Page 257: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

237

this decrease, the peak at 8980 eV which is also a feature of Cu foil spectrum showed up

and increased with time, suggesting the formation of Cu metallic phase. Basing on the

above results, we conclude that changes in the local structure of Cu occurred throughout

the reaction. The quantitative information on the local structure of Cu during the reaction

conditions was obtained by performing EXAFS analysis and the results were summarized

in Figure 6.15. Figure 6.15 shows the change in the coordination numbers of Cu-Cu, Cu-

Mg, and Cu-O bonds during the reaction. From 200-400 °C, a steady decrease in Cu-O

bond coordination number takes place, which, as discussed above, is also manifested by

the increase in the intensity of 8982 eV peak. There was no appearance of Cu-Cu bond

until the steady-state condition at 400 °C. At 400 °C, the final EXAFS spectra show a

significant increase of Cu-Cu coordination number from 0 to about 3. This indicates

clustering of the Cu atoms after reaction has stabilized at 400 °C.

To confirm the correlation between the XANES features with the coordination

number of Cu-O bond, XANES spectra simulations were performed using FEFF 9 code.80

Figure 6.15. Coordination number changes during reaction of ethanol to 1,3-BD over

CuMgSi

Page 258: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

238

Simulations were first performed on CuO and Cu2O to find optimized simulation

parameters, which were then applied in calculating the spectra of all models. For the as-

prepared CuMgSi catalyst, according to EXAFS analysis, the coordination number of Cu-

O was close to 6 and Cu is very likely taking the Mg sites in MgO lattice. We therefore

built an MgO sphere which contains 251 atoms and has diameter of about 1.6 nm, and

replaced the core Mg atom by a Cu atom. This model was named Model 1. In this model,

Cu is octahedrally coordinated by 6 O atoms at the same distance. The calculated XANES

spectrum of this model is plotted in Figure 6.16, and the shoulder peak at the rising edge

is indeed shifted to higher energy compared to that of CuO, which agrees with the trend

observed in experimental data. As shown by EXAFS results, under reaction conditions and

at high temperatures, the average Cu-O decrease and is close to 4. We thus modified Model

1 by removing 2 oxygen atoms around Cu. In this modified model, Model 2, Cu is then

surrounded by 4 oxygen atoms at the same distance forming a planar geometry. In the

simulated XANES spectrum of Model 2, a shoulder peak appears in position between those

of Cu2O and CuO. Such trend was also observed in the experimental spectra. Therefore,

the agreement between the experimental and theoretical XANES spectra suggests the

shoulder peak at the rising edge of Cu spectra is related to the local oxygen environment

around Cu. In the CuMgSi system, Cu replaces Mg in MgO lattice. When the reaction

occurs, the octahedral Cu-O geometry will be distorted: most likely, part of oxygen atoms

are pulling away from Cu, which could be then transformed to Cu metallic phase as

detected in the final aged catalyst (Figure 6.14).

Page 259: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

239

A complementary view of this operando measurement was offered by Angelici et

al., where reactions were carried out at 400 °C under two different pretreatment conditions,

i.e. inert flow and reducing atmosphere.28 Under inert flow, the initial state of the catalyst

consists of the native distorted octahedral Cu2+ species that was originally in the catalyst

and another Cu2+ species that resembles to Cu2+ from CuO/SiO2. This latter Cu2+ species

was reduced to Cu0 and transformed to the distorted octahedral Cu2+ species when

pretreated at 425 °C under inert flow. Our observations show that there are new Cu species

as evident by the peak at 8982 eV that appeared when catalyst was pretreated at high

temperature even though the pre-edge feature at 8977 eV, assigned to the distorted

octahedral Cu2+ from CuMgSi, barely changed. Interestingly, similar distribution between

Cu2+, Cu+, and Cu0 was observed after ethanol reaction without reducing pretreatment, after

Figure 6.16. XANES spectra of the simulated CuO Model 1: Cu in a local environment

surrounded by 6 oxygen atoms and Model 2: Cu in a local environment surrounded by

4 oxygen atoms.

Page 260: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

240

reducing pretreatment under H2 and after ethanol reaction with reducing pretreatment.28

Specifically, the three steps mentioned correspond to increasing amount of Cu0 in the final

state of the catalyst. This indicates that both ethanol and hydrogen have a competing

reducing effect on the catalyst. The final state after the steady-state reaction under both

pretreatment conditions revealed that there were some Cu2+ species on the catalyst even

after extensive reaction with ethanol.28 In our experiments, however, we observed a

different outcome. The two pre-edge features at 8977 and 8987 eV behaved similarly with

both of them barely changing during the reaction. Even after extensive reaction at 400°C,

Cu-Mg coordination number did not change, while Cu-O coordination number decreased

(Figure 6.15) to 4. The apparent increase in peak at 8987 eV is mostly due to the increase

in peak at 8982 eV. We propose, based on data in Figure 6.13-15, that origin of the peak

at 8982 eV, assigned to Cu2+ with less-than-6 oxygen neighbors, is from a bulk Cu2+ with

six oxygen neighbors that catalyzed the reduction and lost bonding with two neighbor

oxygens during interaction with ethanol, as indicated by the simulation (Figure 6.16).

Furthermore, this new Cu species undergoes change in coordination number, decreasing to

reduced Cu0, possibly due to the depleted reducible Cu2+ that shifts the reaction active sites

and further reduced all reducible copper species into Cu0, as suggested by clustering of Cu

(increase in Cu-Cu coordination number) as the reaction progressed at 400°C. The

decreased reducibility of Cu2+, evident from the presence of Cu2+ at the end of the reaction,

was also observed previously on CuZn catalysts supported on MCM-41 and Al2O3, where

co-presence of Zn2+ led to the formation of isolated Cu2+ species that was reduced at higher

temperature.42,81 Other factors that deteriorate Cu2+ reducibility can be attributed to the

presence of solid solution phase and bulk CuO phase, such as that found in CuMnZrO2 and

Page 261: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

241

CuMgAlOx hydrotalcite catalysts, respectively.44,78 The operando XANES and in-situ UV-

Vis confirmed the presence of two Cu species on the catalyst prior to exposure to ethanol,

i.e. distorted octahedral Cu2+ (possibly from solid solution) and reducible Cu2+ species, as

suggested by in-situ dehydrated UV-Vis as well.

3.3.3.2. Operando XANES and EXAFS of ethanol over Zn-promoted MgSi catalyst

The XANES spectra of Zn catalysts and standards taken in ambient condition are

shown in Figure 6.17a. The standards used in this study are Zn foil and ZnO to represent

the reduced and oxidized states of the transition metal. Comparison between ZnMgSi, ZnSi

(ZnO/SiO2), and ZnMg (ZnO/MgO) reveals similarity between ZnMgSi and ZnMg. The

silica-supported sample looks like those of willemite or hemimorphite, both Zn-silicates.47

Chouillet et al. investigated the effect of drying temperature prior to calcination, and

XANES spectra of all dried samples calcined at 450 °C, only 50 °C lower than our

temperature, are nearly identical and indicative of zinc silicate formation.47 The Zn foil

exhibits a peak at 9660 eV, which was assigned to electron transition to empty d orbital.

The absence of this feature indicates that all samples are fully oxidized.55 For Zn standards

(ZnO and Zn foil), there are two main features, the main edge, labeled as A, and feature B

in the spectra. The main peak was assigned to 1s4p electron transition with lesser peak

intensity corresponding to decreasing coordination number of the cation.82–84 The second

feature was a multiple scattering resonance associated with medium range molecular

structure around the target element; this feature was located differently for each sample,

indicating difference in geometric molecular structure.82,83

Page 262: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

242

Both Mg-containing samples, i.e. ZnMg and ZnMgSi, exhibit splitting at the edge

that was significantly larger than that of ZnSi. The splitting was previously observed on

ZnO/Al2O3 and ZnFe2O4 as well and was attributed to a Zn2+ structure in a rigid

environment nothing like ZnO.83,85 EXAFS spectra of the samples show very similar

spectral shape between the two samples although the oscillation magnitude of the ZnMgSi

sample was much lower. The similarity indicates that the Zn in both samples possess very

similar local structure. Fourier transform was applied to the EXAFS signal (k2χ(k)) of

ZnMg to represent both samples and compared to ZnO and Zn foil (Figure 6.17b).

Between 1-3 Å, there are two peaks at 1.40 Å and 2.40 Å. From the Fourier transformed

spectra the first peak was attributed to Zn-O bond, while the latter was lower than Zn-Zn

bond length in ZnO yet higher than Zn-Zn bond length in Zn foil. This implies that this

was not due to the contribution of Zn-Zn bond and we predict this to be Zn-Mg bond. To

confirm it, we did EXAFS analysis for the ZnMgSi catalyst and tested three models: Model

A includes Zn-O and Zn-Zn paths; Model B includes Zn-O, Zn-Zn, and Zn-Mg paths;

Figure 6.17. (a) Normalized XANES spectra of ZnMg, ZnSi, ZnMgSi, Zn foil, and

ZnO. Inset: Zn K-edge k2-weighted EXAFS data of corresponding spectra. (b) Fourier

transforms of the EXAFS spectra of ZnMg, ZnO, and Zn foil.

Page 263: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

243

Model C includes Zn-O and Zn-Mg paths. The fitting k range is 2.0-10.5 Å-1, and R range

is 1.0-3.2 Å. Model 3 provides us best fitting results, which confirms that Zn was singly

distributed into MgO lattice. This Zn-Mg bond was ~0.2 Å shorter than that of Zn-Zn bond

in the ZnO foil, which was also previously determined in Zn(1-x)MgxO alloy.86 The bond

length values for standards and samples are tabulated in Table 6.6.

Figure 6.18. Normalized temperature-programmed operando XANES spectra of

ZnMgSi catalyst under He flow (a) and ethanol flow (b). Inset: enlarged region of the

pre-edge features to elucidate changes at different temperature. (c) Temperature-

induced change in coordination number of Zn-Mg and Zn-O bonds during the reaction.

(c)

Page 264: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

244

The operando XANES spectra during ethanol conversion are presented in Figure

6.18. Similar to study on CuMgSi, the experiment was conducted with increasing

temperature under ethanol flow (Figure 6.18b). The MS data of the experiment shows

similarities with CuMgSi. In particular, acetaldehyde was produced very early as well

following the induction time between ethanol flowing into the reactor and the product

stream entering the MS. The production of 1,3-BD follows similar trend, i.e. started being

produced at lower temperature before really ramping up at ~300 °C. This sudden increase

at 300 °C coincides with the increase in acetaldehyde production as well, which suggests

that there are two active sites for ethanol dehydrogenation for both catalysts. The presence

of these two sites on two promoted catalysts indicates that there are identical sites on both

catalysts. When compared to unpromoted MgSi catalyst, the acetaldehyde production was

found to dramatically increase at this temperature as well. This indicates that Zn and Cu

both are present as an additional dehydrogenating site, and that the native weak basic sites

responsible for the reaction are still present after promotion.

The Zn2+ local structure, however, shows a resilience nature with flowing ethanol,

as shown in Figure 6.18b (inset). There was no significant change under ethanol flow,

compared to the thermal effect when only helium was flown (Figure 6.18a). Figure 6.18c

further shows the analysis of the EXAFS spectra where there was no significant changes

in Zn local coordination number (N) during the reaction. The identified Zn-Mg and Zn-O

both remained intact with no change in the local state of the catalyst was observed. This

indicates that the Zn-promoted catalyst should be relatively stable compared to Cu-

promoted catalyst and possible deactivation is more likely to be related to the formation of

Page 265: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

245

carbonaceous deposit on the surface due to the higher activity exhibited by the additional

redox and Lewis acid sites provided by the Zn dopant.15

Table 6.6. Best fitting results for ZnMgSi, ZnMg, ZnO, MgO and Zn. The structural

parameters of standards were listed for comparison.

Sample Bond N R (Å)

ZnMgSi Zn-O 3.6±0.5 1.98±0.02

Zn-Mg 4.8±1.6 3.09±0.04

ZnMg Zn-O 4.7±1.0 2.09±0.04

Zn-Mg 14.0±2.8 3.05±0.02

ZnO

Zn-O 4 1.94

Zn-Zn 6 3.15

Zn-Zn 6 3.2

MgO Mg-O 6 2.11

Mg-Mg 12 2.98

Zn foil Zn-Zn 6 2.66

Zn-Zn 6 2.88

4. Conclusions

Cu- and Zn-promoted wet kneaded MgO/SiO2 catalysts were interrogated in situ

and operando and provided new insights into the structure and reactivity of their catalytic

sites during ethanol reaction to 1,3-BD. No distinct crystalline promoter phases were

obtained according to XRD and STEM measurements and Cu and Zn was suggested to

bind strongly with the native OH groups. Under dehydrated conditions, oligomeric Cu-O

species were found to dominate CuMgSi while the combination of very small <4 nm ZnO

nanoparticles and possibly solid Zn solution with MgO have been observed using a

combination of UV-Vis and STEM measurements. The reduced amount of strong basic

sites due to the metal promoter binding was found to affect RDS shift from acetaldehyde

formation to MPV reduction. In situ DRIFT spectroscopy results allowed to decouple the

Page 266: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

246

aldol condensation and dehydrogenation fundamental steps that takes place on the surface

suggesting that promoting the catalyst with transition metal promoters enhanced the ability

of the catalyst to carry out aldol condensation as correlated with the steady state reactivity

experiments. In situ UV-Vis spectroscopy suggested appearance of π- π* electronic

transitions of allylic cations, cyclic or aromatic species on the catalysts while also

providing insights on the oligomeric structure of the active sites. In particular, oligomeric

CuO species with ~0.8 Cu nearest neighbor were found to decrease in intensity suggesting

their involvement in ultimate catalytic Cu0 species formation.

Our operando X-ray measurements were combined with ab initio multiple

scattering modelling to unravel the exact electronic structure of the Cu and Zn promoters.

These measurements were performed as a function of temperature and signified that Cu-

Cu bond appeared at reaction temperatures of 400 oC on the aged (TOS of 6-7 hours)

catalyst at the expense of Cu-O bonds. Cu replaced Mg in MgO lattice to eventually lead

to Cu aggregates. This is akin to the literature reports where deactivation of Cu-containing

catalysts was suggested due to the carbonaceous deposits rather than sintering of the

promoter. Furthermore, the 8982 eV peak typically assigned to Cu+ species, in our work

was assigned to a 4-fold coordinate Cu species, rather than Cu2O and is proposed as the

key intermediate leading to increase in Cu-Cu bond number. It is transient and is only

populated at temperatures lower than 400 oC and starts decreasing to yield Cu0 during aging

with ethanol. Two types of Zn bonds, namely Zn-O and Zn-Mg, were identified during X-

ray analysis and showed resilience to ethanol under operating conditions. Particularly, Zn

was nearly 6-coordinated when in the vicinity of Mg while Zn-O species showed nearly 4

nearest neighbors.

Page 267: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

247

Chapter 7 – Supporting Information

S6.1 Catalyst Characterization

Figure S6.1. XRD patterns of (a) Zn/MgO and (b) Zn/SiO2 at different loadings.

Figure S6.2. XRD patterns of (a) Cu/MgO and (b) Cu/SiO2 at different loadings.

Page 268: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

248

Figure S6.3. In-situ DRIFTS of OH region of dehydrated MgSi catalysts references at

100°C for Cu-promoted (left) and Zn-promoted (right). Spectra are offset for clarity.

Figure S6.4. Tauc plot of CuO (left) and deconvoluted Cu species of CuMgSi catalyst

(right) to determine the edge energy/band gap (E0) for correlation with number of Cu

coordination.

Figure S6.5. In-situ UV-Vis difference spectra of oxidative dehydration of (a)

CuMgSi and (b) ZnMgSi.

(a) (b)

Page 269: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

249

S6.2 Catalyst Acid-Base characterization

The CO2 coflow is shown to inhibit 1,3-BD production on all catalysts while also

increasing the production of acetaldehyde, except for ZnMgSi. CO2 interacts only with

strong basic sites, give its nature as a weak acid, and the change in acetaldehyde and

Figure S6.6. Poisoning reactivity testing using CO2 to determine the role of basic sites

during ethanol conversion to 1,3-BD over (a) MgSi, (b) CuMgSi, and (c) ZnMgSi.

Reactions are carried out at 400 °C, mcat = 0.1 g, pethanol = 2.5 kPa, total flow = 55 ml/min.

All formation rates are normalized to initial 1,3-BD formation rate.

Page 270: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

250

ethylene production suggests the role of weak basic sites and strong basic sites in catalyzing

the dehydrogenation and dehydration steps, respectively. The experiment on metal-

promoted catalysts suggest that promotion with Cu or Zn decreases the strong basic sites

since the ethylene production is not severely affected. The interaction between the catalysts’

surface with CO2 was studied by means of in-situ DRIFTS using CO2 as a probe molecule.

The experiment corroborates the in-situ poisoning experiments, where the amount of strong

basic sites, demonstrated by the peaks assigned to monodentate and polydentate carbonate,

are reduced upon introduction of transition metal promoters.87 Upon introduction of the

promoters, change in the stability of both monodentate and polydentate carbonate is

lowered at higher temperature, with carboxylate species is now formed on the catalyst. CO2

adsorption on transition metal oxides generally yields an additional carboxylate species,

which is stabilized by back-bonding between d-orbitals of the metal ion and π* orbital of

the C=O bond.88 Given its much lower stability, this carboxylate must indicate a weaker

basic site that is present on the catalysts. Peak assignments of the CO2 surface species are

tabulated in Table S1.

Page 271: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

251

Table S6.1. Peak assignments of surface CO2 species identified on MgSi, CuMgSi, and

ZnMgSi catalysts.

Species Vibrational mode Catalysts

MgSi CuMgSi ZnMgSi

Monodentate

carbonate

υas OCO 1506 1500 1500

υs OCO 1441 1440 1440

Bidentate

carbonate

υas OCO 1663 1611 1611

υs OCO 1362 - -

Symmetrical υ OCO 1425 - -

Polydentate

Carbonate

υas OCO 1571 1573 1566

υs OCO 1380 1384 1391

Bicarbonate

υas OCO 1634 1644 1644

υs OCO 1460 1464 1464

δ OH 1280 1290 ~1290

Carboxylate υas OCO N/A 1593 1593

υs OCO N/A 1374 1374

Poisoning using propionic acid shows significant reduction in 1,3-BD production

for all catalysts. The reversibility nature of this change indicates that these catalysts in

general had very little amount of strong basic sites. The experiments demonstrated that the

acetaldehyde formation rate was more significantly affected during propionic cofeeding on

promoted catalysts, i.e. CuMgSi and ZnMgSi. From Figure S6.8, it is apparent that the

basicity was very different for the transition metal-promoted catalysts. Propionic acid was

shown to interact more strongly with the transition metal sites deactivating active sites

more readily, as shown by the degree of retardation it caused during propionic acid

cofeeding.

Coflowing NH3 resulted in poisoning of acid sites of the catalysts, and

mechanistically, these sites are responsible for the dehydration steps that follow aldol

condensation.2 In this work, ethylene and 1,3-BD production were adversely affected in

both ZnMgSi and CuMgSi, as opposed to MgSi. Which indicates the higher availability of

total acid sites on unpromoted MgSi catalyst. The decrease in acetaldehyde production on

Page 272: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

252

CuMgSi is due to competitive adsorption and activation of NH3, since CuO catalysts are

well-known SCR catalysts.65,66 Zn promotion, on the other hand, showed a very strong

dehydrogenation enhancement with acetaldehyde being accumulated on the catalyst and

shifting the RDS to MPV reduction.

Page 273: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

253

Figure S6.7. CO2 Temperature Programmed-DRIFTS on (a) MgSi, (b) CuMgSi, and

(c) ZnMgSi.

(b)

(a)

(c)

Page 274: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

254

Figure S6.8. Poisoning reactivity testing using propionic acid to determine the role of

basic sites during ethanol conversion to 1,3-BD over (a) MgSi, (b) CuMgSi, and (c)

ZnMgSi. Reactions are carried out at 400 °C, mcat = 0.1 g, pethanol = 2.5 kPa, total flow

= 55 ml/min. All formation rates are normalized to initial 1,3-BD formation rate.

Page 275: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

255

The surface acidity of the catalysts was investigated with DRIFTS using NH3 as a

probe molecule. NH3 is the most commonly used probe molecule, due to its small size,

which can penetrate all sites available on the catalyst without being limited by catalyst

Figure S6.9. Poisoning reactivity testing using NH3 to determine the role of acid sites

during ethanol conversion to 1,3-BD over (a) MgSi, (b) CuMgSi, and (c) ZnMgSi.

Reactions are carried out at 400 °C, mcat = 0.1 g, pethanol = 2.5 kPa, total flow = 30 ml/min

(without NH3), 55 ml/min (with NH3). All formation rates are normalized to initial 1,3-

BD formation rate. NH3 desorption spectra on MgSi catalysts at 100°C are shown in

(d).

(d)

Page 276: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

256

geometry.89–91 The Lewis acid sites are discriminated by the bending and stretching modes

of the coordinated NH3, i.e. NH3 unpaired electron donated to metal sites and ammonium

ion formed due to the strong OH acid sites.90 From Figure S6.9d, peaks at 1650, 1615,

and 1590 cm-1 are found in the N-H deformation region. The absence of peak at ~1450 cm-

1 indicates the non-existence Brønsted acid sites, while the aforementioned peaks are

assigned to two different Lewis acid sites. Using the help of DFT, peaks at 1650 and 1615

cm-1 are assigned to asymmetric N-H bending mode of Lewis-bound NH3 species, while

1590 cm-1 is assigned to symmetric N-H bending mode of the Lewis-bound NH3 species.

In particular, 1650 cm-1 and 1590 cm-1 represent the same species that disappeared upon

promotion with transition metal sites; assigned to Mg2+3C from our calculation, while the

corresponding symmetric bending mode should be around ~1580 cm-1 give the split. Note

that the DRIFTS simulation of both open and closed sites does not lead to proper

discrimination of both sites, and hence this technique should not be used to differentiate

both sites (Table S2). The Several other peaks at 1490, 1400, and 1370 cm-1 are associated

with dissociative adsorption of NH3 that takes place at low temperature.89

Table S6.2. DFT simulation of NH3 on MgO slab. Simulation was done using VASP, PBE

functionals on 2x2x1 k-point mesh.

Type Binding site Vibrational mode

δas NH2 δs NH2

Open 4C 1616 1590

3C 1598 1557

Closed 4C 1612 1595

3C 1645 1601

Page 277: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

257

To further characterize the active sites on the MgSi catalysts, operando methanol

DRIFTS-MS was carried out due to its versatility as a probe molecule.92,93 On the basic

sites methanol produces CO2, acidic sites yield dimethyl ether (DME), while redox sites

will form formaldehyde.93 Hence, this operando testing allows further measurement of

changes in catalyst redox properties, not directly available via CO2 and NH3 testing. The

DRIFTS spectra of MgSi catalysts are shown in Figure S6.10. The C-H region (Figure

S6.10, left) is typically used to identify the presence of surface methoxide.92 Upon CH3OH

adsorption on the surface, several peaks showed up at 100 °C, i.e., 2990, 2965, 2920, 2860,

2820, and 2780 cm-1. Methoxy species bounded to SiO2 sites, such as Si-OCH3, are

indicated by the peaks at 2990, 2965 and 2860 cm-1, assigned to υas (CH3), υs (CH3), and

2δs (CH3), respectively.94 The analogs of these peaks for Mg-OCH3 are located at 2920,

Figure S6.10. DRIFTS spectra in the C-H stretching (left) and bending (right) region

of methanol desorption under He flow on unpromoted (top) and promoted (bottom)

catalysts.

Page 278: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

258

2820, and 2780 cm-1.95,96 The presence of the two kinds of adsorbed methoxy species have

previously been observed by Wachs and coworkers.94 In their study, it was shown that

surface methoxide was present as Si- and V- bounded in the case of V2O5/SiO2 catalyst. A

shoulder at 2880 cm-1 is clearly seen on Cu- and Zn-promoted samples, but is lower in

intensity for the unpromoted sample. This peak has previously been assigned to the υs (CH)

of formaldehyde, formed by the readsorption of the redox product.94 The peak at >350 °C,

i.e., 2805 cm-1, can possibly be attributed to surface formaldehyde υ (CH), based on

Busca’s work.97

Methanol typically adsorbs as two kinds of species, as surface methoxy

(dissociative adsorption) and as a molecularly bonded species to Lewis acid site in a minor

amount.94 The asymmetric and symmetric methyl bends of the former are located around

~1450 and ~1430 cm-1, while the second species is characterized by its OH bending at

~1360 cm-1 (Figure S6.10, right). These three peaks can be found on the spectra at low

temperatures, while they disappear at higher temperatures. The adsorbed methoxy species

further dehydrogenate into surface formate via C-H bond breaking on the redox site, and

the basic site will perform another C-H bond scission to make carbonate.93 The bicarbonate

species, as explained before, is characterized by peaks at 1644 and 1464 cm-1. The latter is

overlapped by the methoxide methyl bending mode, but still apparent due to the broadness

of the peak. An additional peak at 1670 cm-1 also occurred in these spectra, as it does for

CO2 adsorption, illustrated in Figure S6.10. At higher temperatures monodentate

carbonate is apparent at 1386 cm-1 and is accompanied by a shoulder at ~1587 cm-1. This

is obscured by the intensity of the bidentate carbonate peak at 1611 cm-1. The presence of

surface formate in this experiment is revealed by the peaks at 1595 and a peak at around

Page 279: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

259

~1340 cm-1, assigned to υa (OCO) and υs (OCO), respectively.97 The prevalence of surface

formate on the unpromoted catalyst further demonstrates the basicity of the catalyst. The

peaks at 1595 and 1340 cm-1 are much less pronounced in the spectra of the Cu- and Zn-

promoted catalysts. This could indicate the spontaneous desorption of the produced

formaldehyde, even though at higher temperatures, re-adsorption of formaldehyde is more

pronounced in the case of promoted catalysts, as shown by the peak at 2880 cm-1 (Figure

S6.10, left).

The corresponding MS data from the operando methanol spectroscopy are shown

in Figure S6.11. As discussed, methanol adsorbs in two different ways, by dissociative

adsorption and by molecular adsorption on Lewis sites. This is further corroborated by the

vapor phase MS data, which shows a methanol peak (m/z = 31) for each catalyst, consisting

of two different peaks. The symmetry of these peaks indicates that they consist of two

peaks. A second peak, which is apparent as a shoulder at 300 °C, indicates the release of

two adsorbed methanol species into the vapor-phase. The lower temperature peaks for the

methanol occur at temperatures below 200 °C for each catalyst and are due to the strongly

bound, yet molecularly adsorbed, methanol species on the Lewis acid site. The higher

temperature peak is due to the recombination of the surface methoxide and surface

hydroxyl group. The most striking observation is for the formaldehyde spectra (m/z=29),

where the Tp values are situated close to their corresponding methanol Tp peaks. On redox

sites, methanol dissociates to give surface methoxide (CH3O·) and surface hydroxyl group

(OH·). The surface methoxide will then perform a subsequent C-H bond breaking step and

desorb as formaldehyde.98 The un-promoted sample shows a very close Tp for both

methanol and formaldehyde (190-193 °C), while the Cu-promoted and Zn-promoted

Page 280: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

260

samples show a significantly lowered temperature peak at 177°C. This shows that the redox

capability of the catalyst has been improved by transition metal doping. The formaldehyde

peak at lower temperature is very close to the methanol peak, at 182 and 185 °C for ZnMgSi

and CuMgSi, respectively, which differs only by about 5-8 °C. Formaldehyde is known to

re-adsorb onto the surface94 so the second peak at a higher temperature originates from the

desorption of this formaldehyde species.

The CO2 temperature profiles (m/z=44) are also shown in Figure S6.11. The

mechanism of the methanol to CO2 reaction is complex. Surface formate is required as a

surface intermediate and this requires surface oxygen, since formaldehyde possesses only

one oxygen atom. The re-adsorption of formaldehyde into surface formate is corroborated

by the DRIFTS data shown in Figure S6.10 and induced by the presence of the basic sites

Figure S6.11. Online MS analysis during operando methanol DRIFTS of CuMgSi,

ZnMgSi, MgSi and reference MgO

Page 281: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

261

on the surface.93 Subsequently, C-H bond breaking takes place and then CO2 is released.

Pure MgO is used in this experiment as a reference. On MgO the Tp peaks for methanol,

formaldehyde and CO2 are very close to each other, at 198, 200, and 200 °C, respectively.

This indicates a competing, consecutive reaction for both formaldehyde and CO2 formation.

In this case, the desorption of formaldehyde from the surface after methoxy C-H bond

breaking has a very similar rate with subsequent HCOO- formation and C-H bond breaking

to give off CO2. The redox capability of MgO is also acknowledged by Badlani and Wachs,

where a steady-state reaction of MgO yields both formaldehyde and CO2 with higher

selectivity towards the former.93 The presence of the second CO2 peak at 315 °C represents

the secondary formation of CO2 from the re-adsorption of formaldehyde. The CO2 peak

never fully disappears, since there is a small amount of CO2 being released from bulk

magnesium carbonate.99 Interestingly, this basicity of MgO is not reflected in the MgSi

catalyst. The CO2 peak is practically non-apparent as shown in Figure S6.11. SiO2 is a very

inert material, which should be the reason why there is very little CO2 in the vapor-phase.

This is mostly due to methanol thermal decomposition.93 Wet-kneading with SiO2 (1:1)

should reduce the number of MgO basic sites and induce the formation of different sites,

such as closed or open Lewis acid sites. These are formed from the Mg-O-Si linkages11

and additional redox sites, as shown by the increased formaldehyde production for MgSi.

The increased number of redox sites is apparently caused moreso by the higher

formaldehyde productivity than by the pure MgO, which is why most MgO catalysts won’t

perform the ethanol-to-1,3-BD reaction. From our CO2 DRIFTS experiment in Figure S6.7,

it can be seen that the CO2 peaks are present on all catalysts at elevated temperatures, with

MgSi possessing the most integrated area. Un-promoted MgSi shows a better retention of

Page 282: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

262

CO2, demonstrated by the higher intensity of the carbonate peaks at a higher temperature,

as compared to ZnMgSi and CuMgSi (Figure S6.7). It is likely that once the C-H bond of

the surface formate is broken, the CO2 formed is bound to an oxygen site on the

unpromoted catalyst, whereas promotion with Zn or Cu reduced the electronegativity of

the nearby oxygen site and released CO2 at a higher rate than for the unpromoted sample.

A critical analysis done in this work is the comparison of the redox capability of

the catalysts, which was done by evaluating the activation energy of the methanol

dehydrogenation reaction to yield formaldehyde. The decomposition kinetics of the C-H

bond breaking of surface methoxy is known to follow a first order reaction with pre-

exponential factor of ~1013 s-1.100 The activation energy was calculated using the Redhead

equation101

𝐸𝑎

𝑅𝑇𝑝2 =

𝜐

𝛽𝑒

(−𝐸𝑎

𝑅𝑇𝑝) (1)

where Ea is the activation energy (J/mol), R is the gas constant (J/mol K), Tp is the TPSR

temperature (K), β is the heating rate (10 °C/min), and υ is the pre-exponential factor (s-1).

From equation (1), activation energies of 28.4, 27.7 and 26.7 kcal/mol for reference

MgO, MgSi and the CuMgSi and ZnMgSi were calculated, respectively. The lowered

activation energy explains the more reactive nature of the catalyst, since promotion of the

catalyst has been shown to give a lower activation energy for the dehydrogenation reaction.

Promotion with Zn and Cu enhances the alcohol dehydrogenation capability.13,14 This

promotional effect doesn’t carry over to the ethanol dehydrogenation in a straightforward

manner, since for ethanol, utilizing Zn and Cu promotion has shown a very profound effect

on the ethanol dehydrogenation step of the reaction.3 Similarly, the non-existence of DME

Page 283: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

263

as the product of the acidic sites can’t be translated to the catalyst inability to carry out the

dehydration reaction. DME formation requires two available sites that bind two ethoxy

species, while for ethylene formation only one site is necessary. The absence of DME in

the product goes along well with our steady-state experiment, where hardly any diethyl

ether (DEE), an analog for ethanol bimolecular dehydration, is produced. The inability of

the catalyst to produce DEE does not necessarily mean it is not acidic, since ethylene is

prominently produced during the reaction. This experiment suggests that all the catalysts

used in this work do not possess two neighboring acidic sites, which are required to

dehydrate alcohols. Furthermore, the semi-quantification of the active sites for redox sites

can be calculated by integrating the area under the peak, referenced to the surface area of

the catalysts, which are calculated using BET method (Table S6.3).

Table S6.3. Redox properties of the MgSi, CuMgSi and ZnMgSi catalysts and reference

MgO obtained from MS measurements. These results have been normalized to the BET

surface area (m2/g) of each catalysts.

Catalysts Redox site density relative

to MgSi

Activation energy

(kcal/mol)

Reference MgO 0.48 28.4

MgSi 1.00 27.7

CuMgSi 0.42 26.7

ZnMgSi 0.96 26.7

Promotion of the catalyst with a transition metal unexpectedly reduced the number

of redox sites, which indicates the loss of these sites when the catalyst is promoted with

transition metals. Promotion with Cu reduced the Redox site density significantly, while

Zn barely modified the redox site density, which can be related to the superior ethanol

reactivity in the steady-state experiment. The decrease in Redox site density further

Page 284: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

264

justifies explanations by previous reports that higher loading would be detrimental to the

catalyst activity, as well as Zn’s significant enhancement in the ethanol’s conversion.3

S6.3 In-situ and operando characterization of Cu and Zn coordination on promoted

MgSi catalysts

Figure S6.12. In-situ UV-Vis DRS of ethanol reaction on undoped MgO/SiO2 catalyst.

Difference spectra is shown, where catalyst spectra at 100°C with chemisorbed ethanol is

used as a reference.

Figure S6.13. R-space EXAFS spectra of CuMg catalyst, in comparison to Cu foil, CuO,

and Cu2O

Page 285: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

265

Figure S6.14. Corresponding MS data of in-situ XANES-EXAFS for ethanol to 1,3-

BD over (a) CuMgSi, (b) ZnMgSi

Page 286: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

266

References

(1) Posada, J. A.; Patel, A. D.; Roes, A.; Blok, K.; Faaij, A. P. C.; Patel, M. K.

Bioresour. Technol. 2013, 135, 490–499.

(2) Shylesh, S.; Gokhale, A. A.; Scown, C. D.; Kim, D.; Ho, C. R.; Bell, A. T.

ChemSusChem 2016, 9 (12), 1462–1472.

(3) Pomalaza, G.; Capron, M.; Ordomsky, V.; Dumeignil, F. Catalysts 2016, 6 (12),

203.

(4) Sushkevich, V. L.; Ivanova, I. I. ChemSusChem 2016, 9 (16), 2216–2225.

(5) Sushkevich, V. L.; Palagin, D.; Ivanova, I. I. ACS Catal. 2015, 4833–4836.

(6) Taifan, W. E.; Bučko, T.; Baltrusaitis, J. J. Catal. 2017, 346, 78–91.

(7) Muller, P.; Burt, S. P.; Love, A. M.; McDermott, W. P.; Wolf, P.; Hermans, I. ACS

Catal. 2016, 6 (10), 6823–6832.

(8) Sushkevich, V. L.; Ivanova, I. I. Appl. Catal. B Environ. 2017, 215, 36–49.

(9) Chieregato, A.; Velasquez Ochoa, J.; Bandinelli, C.; Fornasari, G.; Cavani, F.;

Mella, M. ChemSusChem 2014, 8 (2), 377–388.

(10) Zhang, M.; Gao, M.; Chen, J.; Yu, Y. RSC Adv. 2015, 5 (33), 25959–25966.

(11) Taifan, W.; Yan, G. X.; Baltrusaitis, J. Catal. Sci. Technol. 2017, 7 (20), 4648–

4668.

(12) Fan, D.; Dong, X.; Yu, Y.; Zhang, M. Phys. Chem. Chem. Phys. 2017, 19 (37),

25671–25682.

(13) Hayashi, Y.; Akiyama, S.; Miyaji, A.; Sekiguchi, Y.; Sakamoto, Y.; Shiga, A.;

Koyama, T.; Motokura, K.; Baba, T. Phys. Chem. Chem. Phys. 2016, 18 (36),

25191–25209.

(14) Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A.

ChemSusChem 2014, 7 (9), 2505–2515.

(15) Larina, O.; Kyriienko, P.; Soloviev, S. Catal. Letters 2015, 1–7.

(16) Da Ros, S.; Jones, M. D.; Mattia, D.; Pinto, J. C.; Schwaab, M.; Noronha, F. B.;

Kondrat, S. A.; Clarke, T. C.; Taylor, S. H. ChemCatChem 2016, 8, 1–12.

(17) Ohnishi, R.; Akimoto, T.; Tanabe, K. J. Chem. Soc. Chem. Commun. 1985, No.

22, 1613–1614.

(18) Sushkevich, V. L.; Ivanova, I. I.; Ordomsky, V. V; Taarning, E. ChemSusChem

2014, 7 (9), 2527–2536.

(19) Makshina, E. V.; Janssens, W.; Sels, B. F.; Jacobs, P. A. Catal. Today 2012, 198

(1), 338–344.

(20) Jones, M. D.; Keir, C. G.; Iulio, C. Di; Robertson, R. A. M.; Williams, C. V;

Apperley, D. C. Catal. Sci. Technol. 2011, 1 (2), 267–272.

(21) Bond, G. C.; Thompson, D. T. Catal. Rev. - Sci. Eng. 1999, 41 (3 & 4), 319–388.

(22) Guan, Y.; Hensen, E. J. M. J. M. Appl. Catal. A Gen. 2009, 361 (1), 49–56.

(23) Wittcoff, H. A. A. J. Chem. Educ. 1983, 60 (12), 1044.

(24) Sushkevich, V. L.; Ivanova, I. I.; Taarning, E. ChemCatChem 2013, 5 (8), 2367–

2373.

(25) Chang, F.-W.; Kuo, W.-Y.; Lee, K.-C. Appl. Catal. A Gen. 2003, 246 (2), 253–

264.

(26) Chang, F.-W.; Yang, H.-C.; Roselin, L. S.; Kuo, W.-Y. Appl. Catal. A Gen. 2006,

304 (0), 30–39.

Page 287: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

267

(27) Janssens, W.; Makshina, E. V. V.; Vanelderen, P.; De Clippel, F.; Houthoofd, K.;

Kerkhofs, S.; Martens, J. A. A.; Jacobs, P. A. A.; Sels, B. F. B. F. ChemSusChem

2014, 8 (6), 994–1008.

(28) Angelici, C.; Meirer, F.; van der Eerden, A. M. J.; Schaink, H. L.; Goryachev, A.;

Hofmann, J. P.; Hensen, E. J. M.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ACS

Catal. 2015, 5 (10), 6005–6015.

(29) Kyriienko, P. I.; Larina, O. V; Soloviev, S. O.; Orlyk, S. M.; Calers, C.; Dzwigaj,

S.; Pisarzhevsky, L. V. ACS Sustain. Chem. Eng. 2017, 5 (3), 2075–2083.

(30) Sekiguchi, Y.; Akiyama, S.; Urakawa, W.; Koyama, T.; Miyaji, A.; Motokura, K.;

Baba, T. Catal. Commun. 2015, 68, 20–24.

(31) Baylon, R. a. L.; Sun, J.; Wang, Y. Catal. Today 2015, 1–7.

(32) Musolino, V.; Selloni, A.; Car, R. J. Chem. Phys. 1998, 108 (12), 5044.

(33) Pacchioni, G. J. Chem. Phys. 1996, 104 (18), 7329.

(34) Asakura, K.; Iwasawa, Y. Mater. Chem. Phys. 1988, 18, 499–512.

(35) Colonna, S.; Arciprete, F.; Balzarotti, A.; Fanfoni, M.; De Crescenzi, M.; Mobilio,

S. Surf. Sci. 2002, 512, L341–L345.

(36) Rodriguez, J. a.; Jirsak, T.; Chaturvedi, S. J. Chem. Phys. 1999, 111 (17), 8077.

(37) Rodriguez, J. a.; Jirsak, T.; Freitag, A.; Larese, J. Z. J. Phys. Chem. B 2000, 2

(100), 7439–7448.

(38) Zhukovskii, Y. F.; Kotomin, E. a.; Borstel, G. Vacuum 2004, 74 (2), 235–240.

(39) Matveev, A.; Neyman, K.; Yudanov, I.; Rösch, N. Surf. Sci. 1999, 426 (1), 123–

139.

(40) Galadima, A.; Muraza, O. J. Nat. Gas Sci. Eng. 2015, 25, 303–316.

(41) Zonetti, P. C. C.; Celnik, J.; Letichevsky, S.; Gaspar, A. B. B.; Appel, L. G. G. J.

Mol. Catal. A Chem. 2011, 334 (1–2), 29–34.

(42) Velu, S.; Suzuki, K.; Okazaki, M.; Kapoor, M. P.; Osaki, T.; Ohashi, F. J. Catal.

2000, 194 (2), 373–384.

(43) Ro, I.; Liu, Y.; Ball, M. R. R.; Jackson, D. H. K. H. K.; Chada, J. P. P.; Sener, C.;

Kuech, T. F. F.; Madon, R. J. J.; Huber, G. W. W.; Dumesic, J. A. A. ACS Catal.

2016, 6 (10), 7040–7050.

(44) Bravo-Suarez, J. J.; Subramaniam, B.; Chaudhari, R. V. J. Phys. Chem. C 2012,

116 (34), 18207–18221.

(45) Vlaic, G.; Bart, J. C. J. C. J.; Cavigiolo, W.; Mobilo, S. Chem. Phys. Lett. 1980, 76

(3), 453–459.

(46) Witzke, M. E. E.; Dietrich, P. J. J.; Ibrahim, M. Y. S. Y. S.; Al-Bardan, K.;

Triezenberg, M. D. D.; Flaherty, D. W. W. Chem. Commun. 2017, 53 (3), 597–

600.

(47) Chouillet, C.; Villain, F.; Kermarec, M.; Lauron-Pernot, H.; Louis, C. J. Phys.

Chem. B 2003, 107 (15), 3565–3575.

(48) Didenko, O. Z.; Kosmambetova, G. R.; Strizhak, P. E. Chinese J. Catal. 2008, 29

(11), 1079–1083.

(49) Pascual, J. L.; Savoini, B.; González, R. Phys. Rev. B 2004, 70 (4), 45109.

(50) Hosoglu, F.; Faye, J.; Mareseanu, K.; Tesquet, G.; Miquel, P.; Capron, M.;

Gardoll, O.; Lamonier, J. F.; Lamonier, C.; Dumeignil, F. Appl. Catal. A Gen.

2015, 504, 533–541.

(51) Laskowski, R.; Blaha, P.; Schwarz, K. Phys. Rev. B 2003, 67 (7), 75102.

Page 288: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

268

(52) Breedon, M.; Spencer, M. J. S.; Yarovsky, I. Surf. Sci. 2009, 603 (24), 3389–3399.

(53) Gruver, V.; Sun, A.; Fripiat, J. J. Catal. Letters 1995, 34 (3–4), 359–364.

(54) Chung, S.-H.; Angelici, C.; Hinterding, S. O. M.; Weingarth, M.; Baldus, M.;

Houben, K.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ACS Catal. 2016, 6 (6),

4034–4045.

(55) Harvey, J. N. Kaltsoyannis, N., McGrady, J. E., Eds.; Springer Berlin Heidelberg:

Berlin, Heidelberg, Heidelberg, 2004; pp 151–184.

(56) Chizallet, C.; Costentin, G.; Che, M.; Delbecq, F.; Sautet, P.; Surface, D.; Marie,

P. J. Am. Chem. Soc. 2007, No. 19, 6442–6452.

(57) Gawande, M. B.; Goswami, A.; Felpin, F.-X.; Asefa, T.; Huang, X.; Silva, R.;

Zou, X.; Zboril, R.; Varma, R. S. Chem. Rev. 2016, 116 (6), 3722–3811.

(58) Marion, M. C.; Garbowski, E.; Primet, M. J. Chem. Soc.{,} Faraday Trans. 1990,

86 (17), 3027–3032.

(59) Coluccia, S.; Lavagnino, S.; Marchese, L. Mater. Chem. Phys. 1988, 18 (5), 445–

464.

(60) Yoshida, H.; Shimizu, T.; Murata, C.; Hattori, T. J. Catal. 2003, 220 (1), 226–232.

(61) Cavalleri, M.; Pelmenschikov, A.; Morosi, G.; Gamba, A.; Coluccia, S.; Martra, G.

In Oxide-based Systems at the Crossroads of ChemistrySecond International

Workshop October 8-11, 2000, Como, Italy; A. Gamba, C. C. and S. C. B. T.-S. in

S. S. and C., Ed.; Elsevier, 2001; Vol. Volume 140, pp 131–139.

(62) Valizadeh, H.; Azimi, A. A. J. Iran. Chem. Soc. 2011, 8 (1), 123–130.

(63) De Baerdemaeker, T.; Feyen, M.; Müller, U.; Yilmaz, B.; Xiao, F.-S.; Zhang, W.;

Yokoi, T.; Bao, X.; Gies, H.; De Vos, D. E. ACS Catal. 2015, 3393–3397.

(64) Ho, C. R.; Shylesh, S.; Bell, A. T. ACS Catal. 2016, 6 (2), 939–948.

(65) Liu, Q.; Liu, Z.; Li, C. Chinese J. Catal. 2006, 27 (7), 636–646.

(66) Díaz, G.; Pérez-Hernández, R.; Gómez-Cortés, A.; Benaissa, M.; Mariscal, R.;

Fierro, J. L. G. J. Catal. 1999, 187 (1), 1–14.

(67) Baltrusaitis, J.; Mendoza-Sanchez, B.; Fernandez, V.; Veenstra, R.; Dukstiene, N.;

Roberts, A.; Fairley, N. Appl. Surf. Sci. 2015, 326, 151–161.

(68) Georgieff, K. K. J. Appl. Polym. Sci. 1966, 10 (9), 1305–1313.

(69) Bevington, J. C.; Norrish, R. G. W. Proc. R. Soc. London. Ser. A. Math. Phys. Sci.

1949, 196 (1046), 363 LP-378.

(70) Song, H.; Ozkan, U. S. J. Catal. 2009, 261 (1), 66–74.

(71) Inui, K.; Kurabayashi, T.; Sato, S. Appl. Catal. A Gen. 2002, 237 (1–2), 53–61.

(72) Singh, M.; Zhou, N.; Paul, D. K.; Klabunde, K. J. J. Catal. 2008, 260 (2), 371–

379.

(73) Palagin, D.; Sushkevich, V. L.; Ivanova, I. I. J. Phys. Chem. C 2016, 120 (41),

23566–23575.

(74) Ordomsky, V. V.; Sushkevich, V. L.; Ivanova, I. I. J. Mol. Catal. A Chem. 2010,

333 (1), 85–93.

(75) Wulfers, M. J.; Tzolova-Müller, G.; Villegas, J. I.; Murzin, D. Y.; Jentoft, F. C. J.

Catal. 2012, 296, 132–142.

(76) Nordvang, E. C.; Borodina, E.; Ruiz-Martínez, J.; Fehrmann, R.; Weckhuysen, B.

M. Chem. – A Eur. J. 2015, 21 (48), 17324–17335.

(77) Kang, M.; Park, E. D.; Kim, J. M.; Yie, J. E. Catal. Today 2006, 111 (3–4), 236–

241.

Page 289: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

269

(78) Xu, R.; Wei, W.; Li, W. H.; Hu, T. D.; Sun, Y. H. J. Mol. Catal. A Chem. 2005,

234 (1–2), 75–83.

(79) Choy, J.-H.; Kim, D.-K.; Hwang, S.-H.; Demazeau, G. Phys. Rev. B 1994, 50 (22),

16631–16639.

(80) Rehr, J. J.; Kas, J. J.; Vila, F. D.; Prange, M. P.; Jorissen, K. Phys. Chem. Chem.

Phys. 2010, 12 (21), 5503–5513.

(81) Velu, S.; Wang, L.; Okazaki, M.; Suzuki, K.; Tomura, S. Microporous

Mesoporous Mater. 2002, 54 (1–2), 113–126.

(82) Galoisy, L.; Cormier, L.; Calas, G.; Briois, V. J. Non. Cryst. Solids 2001, 293–295,

105–111.

(83) Wang, H.-C.; Wei, Y.-L.; Yang, Y.-W.; Lee, J.-F. J. Electron Spectros. Relat.

Phenomena 2005, 144–147, 817–819.

(84) Kelly, R. A.; Andrews, J. C.; DeWitt, J. G. Microchem. J. 2002, 71 (2), 231–245.

(85) Waychunas, G. A.; Fuller, C. C.; Davis, J. A.; Rehr, J. J. Geochim. Cosmochim.

Acta 2003, 67 (5), 1031–1043.

(86) Jeong, E.-S.; Park, C.; Jin, Z.; Yoo, J.; Yi, G.-C.; Han, S.-W. Journal of

Nanoscience and Nanotechnology. pp 1880–1883.

(87) Di Cosimo, J. I.; Dıez, V. K.; Xu, M.; Iglesia, E.; Apesteguıa, C. R. J. Catal. 1998,

178 (2), 499–510.

(88) Busca, G.; Lorenzelli, V. Mater. Chem. 1982, 7 (1), 89–126.

(89) T. Barzetti, E. Selli, D. Moscotti, L. F. J. Chem. Soc. Faraday Trans. 1996, 92 (8),

1401–1407.

(90) Lercher, J. A.; Grijndling, C.; Eder-Mirth, G. Catal. Today 1996, 27, 353–376.

(91) Lavalley, J. C. Catal. Today 1996, 27, 377–401.

(92) Burcham, L. J.; Briand, L. E.; Wachs, I. E. Langmuir 2001, 17 (20), 6164–6174.

(93) Badlani, M.; Wachs, I. E. Catal. Letters 2001, 75 (3–4), 137–149.

(94) Burcham, L. J.; Deo, G.; Gao, X.; Wachs, I. E. Top. Catal. 2000, 11–12, 85–100.

(95) Di Valentin, C.; Del Vitto, A.; Pacchioni, G.; Abbet, S.; Wörz, A. S. S.; Judai, K.;

Heiz, U. J. Phys. Chem. B 2002, 106 (46), 11961–11969.

(96) Kondo, J.; Sakata, Y.; Maruya, K.; Tamaru, K.; Onishi, T. Appl. Surf. Sci. 1987,

28, 475–478.

(97) Busca, G.; Lamotte, J.; Lavalley, J. ciaude; Lorenzelli, V. J. Am. Chem. Soc. 1987,

109 (17), 5197–5202.

(98) Kim, T.; Wachs, I. E. J. Catal. 2008, 255 (2), 197–205.

(99) Angelici, C.; Velthoen, M. E. Z.; Weckhuysen, B. M.; Bruijnincx, P. C. A. Catal.

Sci. Technol. 2015.

(100) Feng, T.; Vohs, J. M. J. Catal. 2002, 208 (2), 301–309.

(101) Redhead, P. a. Vacuum 1962, 12 (4), 203–211.

Page 290: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

270

Chapter 7

Conclusions and future outlook

1. Conclusions ........................................................................................................270

2. Future Outlook ..................................................................................................273

References ......................................................................................................................275

1. Conclusions

Chapter 3

A complex reactive mechanism of ethanol to form 1,3-butadiene was explored

using periodic quantum chemical methods. Three reaction mechanisms, in particular, were

tested using DFT, namely Prins condensation, aldol condensation, and hemiacetal

rearrangements. Based on the thermodynamic and kinetic data determined within this study

we identified four rate important steps in the overall process, namely ethanol

dehydrogenation and dehydration to acetaldehyde and ethylene, respectively, aldol

condensation between adsorbed enolate and physisorbed acetaldehyde, and finally

carbanion stabilization between ethylene and acetaldehyde. In particular, ethanol

dehydration to form ethylene possessed lower energy barrier than dehydrogenation to yield

acetaldehyde suggesting competing reactive pathways. Aldol condensation step to form

acetaldol was preceded with forward free-energy barrier of 16.1 kcal/mol but limited

thermodynamically with endergonic reaction free energy of 12.9 kcal/mol. The energetic

barrier for the first C-C bond formation step in the Prins condensation mechanism at 28.8

kcal/mol also demonstrated the viability of this mechanism over an idealized MgO defect

sites. The model employed here represents a simplified, ideal defected MgO surface,

Page 291: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

271

without taking into account participation of the formed Mg-O-Si bonds resulting from

interaction between MgO-SiO2 and OH groups, significantly affected the acidity-basicity

of the catalyst.

Chapter 4

Surface chemistry of WK (1:1) catalyst during the reaction of ethanol and the

corresponding reactive intermediates, including acetaldehyde, crotonaldehyde, crotyl

alcohol, was investigated using in situ DRIFTS measurements combined with DFT

calculations. Involvement of the native hydroxyl groups was shown to be transient, mostly

due to the hydrogen bonding with the intermediates. The stability of these OH groups

suggested that interaction between the catalyst and intermediates might be due to the

interaction between the Lewis metal heteroatom with the intermediates, instead of with the

OH groups. Ethanol adsorbed as both physisorbed and chemisorbed surface species, while

acetaldehyde, when formed exhibited high reactivity to yield crotonaldehyde but the excess

resulted in strongly bound surface species assigned to surface acetate, and/or 2,4-

hexadienal or polymerized acetaldehyde due to the basicity of the surface. Crotonaldehyde

was more likely to be reduced by ethanol to yield crotyl alcohol than desorbing, even at

relatively high temperatures. DRIFTS study of crotyl alcohol further elucidated the nature

of its interaction with the catalyst, where dissociative adsorption led to the deprotonation

of the molecule and C-O bond scission to yield 1,3-BD. Altogether, the data presented

unraveled a complex interplay between the surface hydroxyl groups, gaseous reactants and

surface bound reactive intermediates of 1,3-BD formation.

Chapter 5

MgO/SiO2 catalyst active surface sites were analyzed using in situ DRIFTS (using

Page 292: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

272

complementary DFT calculations), TPRS and steady state reactor in combination with bulk

XRD and surface LEIS measurements. Combination of in-situ probing with CO2 and

pyridine and in-situ poisoning demonstrated the site requirement for the catalyst. In

particular, it was determined that the weak basic sites were responsible for ethanol

dehydrogenation, strong basic sites for aldol condensation and MPV reduction, while

stronger acid sites catalyzed acetaldol and crotyl alcohol dehydration reactions and weak

acid sites catalyzed the undesired ethanol dehydration. Furthermore, through a combination

of NH3-TPD and DFT the presence of open and closed LAS was identified while further

elaborating Mg coordination, as adopted from LAS classification of zeolitic materials.1–3

The MgSi-WK catalyst was shown to have both open LAS with both Mg3C and Mg4C as

the anchoring LAS, and a very isolated closed LAS (Mg3C).

Chapter 6

Cu- and Zn-promoted wet kneaded MgO/SiO2 catalysts were interrogated in situ

and operando and provided new insights into the structure and reactivity of their catalytic

sites during ethanol reaction to 1,3-BD. In-situ UV-Vis revealed the presence of Cu2+

species with dimeric coordination, while for Zn-promoted MgO/SiO2 catalyst, bulk ZnO

phase and Zn-MgO solid solution were observed. Promotion with metals showed increases

in weak basic sites, with Zn contributing to more Lewis acid sites that are responsible for

the enhanced activity. In-situ DRIFT spectroscopy results allowed decoupling of the aldol

condensation and dehydrogenation fundamental steps that took place on the surface

suggesting that promoting the catalyst with transition metal promoters enhanced the ability

of the catalyst to carry out aldol condensation as correlated with the steady state reactivity

experiments. In-situ UV-Vis spectroscopy suggested appearance of π-π* electronic

Page 293: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

273

transitions of allylic cations, cyclic or aromatic species on the catalysts while also

providing insights on the oligomeric structure of the active sites. Our operando X-ray

measurements were combined with ab initio multiple scattering modelling to unravel the

exact electronic structure of the Cu and Zn promoters. Change in the local coordination of

Cu indicates the presence of more than one Cu2+ species, with only one contributes to the

reaction. This Cu2+ species was reduced to Cu0 during reaction via a Cu species that was

identified as a Cu species with reduced Cu-O coordination number. Zn-promotion, on the

other hand, resulted in a very stable catalyst with stable Zn local coordination, barely

changed during the reaction. However, this catalyst exhibited very high reactivity, which

results in the formation of carbonaceous deposit that further deactivated the catalyst.

2. Future Outlook

The most important issue in the Lebedev reaction is to design a selective catalyst

that possesses an optimum combination of redox, basic, and acid sites. The lack of suitable

spectroscopic methods hampers comprehensive characterization of MgO/SiO2 catalysts.

Up until now, structure-activity relationship has not been achieved yet, with previous

investigators can only indirectly correlate the amount of layered hydrous magnesium

silicate phase to the 1,3-BD yield.4 Work by Hayashi, et al. further proves that SiO2 is not

fundamentally required for this reaction, which suggests that more comprehensive

characterization is necessary to directly correlate the molecular structure of the catalyst that

actively catalyze the reaction.5 The presence of both open and closed Lewis acid sites

discovered in this work further open a new research pathway, where combination of

spectroscopic method and probe molecules is necessary to unravel their participation

Page 294: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

274

during reaction. Further study to confirm the reaction mechanism is also important to

further elucidate the need of specific active sites. Understanding the mechanism of this

cascade reaction, combined with elucidation of the molecular structure of the catalyst will

lead to a more rational design of the catalyst. In particular, the presence of water and

acetaldehyde during the reaction needs to be investigated.6,7

This system is still on an early stage, with no consensus on which preparation

method, Mg/Si ratio, optimum transition metal loading, and calcination temperature

achieved. The synthesis parameter leads to different acidity and basicity of the catalyst,

which suggests the need to optimize these parameters. Optimized parameters will lead to a

superior catalyst with the best selectivity, and combined with knowledge of reaction

mechanism, a kinetic rate expression can be modeled to engineer an appropriate reactive

system.

Page 295: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

275

References

(1) Harris, J. W.; Cordon, M. J.; Di Iorio, J. R.; Vega-Vila, J. C.; Ribeiro, F. H.;

Gounder, R. J. Catal. 2016, 335, 141–154.

(2) Boronat, M.; Concepcion, P.; Corma, A.; Navarro, M. T.; Renz, M.; Valencia, S.

Phys. Chem. Chem. Phys. 2009, 11 (16), 2876–2884.

(3) Boronat, M.; Concepción, P.; Corma, A.; Renz, M.; Valencia, S. J. Catal. 2005,

234 (1), 111–118.

(4) Chung, S.-H.; Angelici, C.; Hinterding, S. O. M.; Weingarth, M.; Baldus, M.;

Houben, K.; Weckhuysen, B. M.; Bruijnincx, P. C. A. ACS Catal. 2016, 6 (6),

4034–4045.

(5) Hayashi, Y.; Akiyama, S.; Miyaji, A.; Sekiguchi, Y.; Sakamoto, Y.; Shiga, A.;

Koyama, T.; Motokura, K.; Baba, T. Phys. Chem. Chem. Phys. 2016, 18 (36),

25191–25209.

(6) Velasquez Ochoa, J.; Bandinelli, C.; Vozniuk, O.; Chieregato, A.; Malmusi, A.;

Recchi, C.; Cavani, F. Green Chem. 2016, 18, 1653–1663.

(7) Zhu, Q.; Wang, B.; Tan, T. ACS Sustain. Chem. Eng. 2017, 5 (1), 722–733.

Page 296: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

276

Curriculum Vitae

William E. Taifan

Address: 9 Duh Drive #131, Bethlehem, PA 18015

Phone: +1 510-219-0716; +62 812-1601-1180

Email: [email protected]

Place of Birth: Surabaya, Jawa Timur, Indonesia

Date of Birth: September 25, 1991

Parents: Fantasi Pola and Ernie Gonawan

Experience

Research Assistant at Lehigh University

August 2014 – present

Ethanol-to-chemicals catalysis research. Presently focused in n-butanol and 1,3-butadiene

synthesis, as well as ethanol as a hydrogenating agent for bio-oil. Side projects including

CO2 capture and theoretical work on oxidative coupling of methane (OCM) catalyst.

Intern at Chemisence, Inc.

February 2014 - June 2014 (5 months)

Research internship at an upcoming green tech startup in Silicon Valley. Research

focused on developing chemical sensors based on carbon black conductivity and affinity

with polymers.

Undergraduate Research Assistant at Biochemical Engineering Lab ITS

August 2012 - August 2013 (1 year 1 month)

Undergraduate research on separation and purification of an ester from a bio-oil.

Technical Expertise

• Matlab, Vienna Ab-initio Simulation Package (VASP) and Gaussian09

• Catalyst synthesis, testing and characterization

• Spectroscopy (IR, UV-Vis)

• GC, GC-MS

Page 297: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

277

Academic Qualifications

Doctor of Philosophy (PhD) in Chemical Engineering

Lehigh University 2014-2018

Dissertation Title:

Advisor: Professor Jonas Baltrusaitis

Master of Science in Chemical Engineering

University of California at Berkeley 2013-2014

Bachelor of Engineering (with Honor)

Institut Teknologi Sepuluh Nopember Surabaya, Indonesia 2009-2013

Publications

“Operando structure determination of Cu and Zn on supported MgO/SiO2 catalysts

during ethanol conversion to 1,3-butadiene”

ACS Catalysis, submitted

Authors: William E. Taifan, Yuanyuan Li, John P. Baltrus, Anatoly I. Frenkel, Lihua

Zhang and Jonas Baltrusaitis

“In-situ spectroscopic insights on the molecular structure of the MgO/SiO2 catalytic

active site during ethanol conversion to 1,3-butadiene”

Journal of Physical Chemistry C, submitted

Authors: William Taifan and Jonas Baltrusaitis

“Surface chemistry of MgO/SiO2 catalysts during the ethanol catalytic conversion to

1,3-butadiene: in situ DRIFTS and DFT study”

Catalysis Science & Technology, 2017, 7(20), 4648-4668

Authors: William E. Taifan, George X. Yan, Jonas Baltrusaitis

“CH4 and H2S reforming to CH3SH and H2 catalyzed by metal promoted Mo6S8

cluster: a first-principles micro-kinetic study”

Catalysis Science & Technology, 2017, 7 (16), 3546-3554

Authors: William E. Taifan, Adam A. Arvidsson, Eric Nelson, Anders Hellman and

Jonas Baltrusaitis

“Minireview: direct catalytic conversion of sour natural gas (CH 4+ H 2 S+ CO 2)

components to high value chemicals and fuels”

Catalysis Science & Technology, 2017, 7 (14), 2919-2929

Authors: William E. Taifan and Jonas Baltrusaitis

“Catalytic conversion of ethanol to 1,3-butadiene on MgO: a comprehensive

mechanism elucidation using DFT calculations”

Page 298: Catalytic Transformation of Ethanol to 1,3-Butadiene over MgO/SiO2 Catalyst

278

Journal of Catalysis, 2017, 346, 78–91

Authors: William E. Taifan, Tomas Bucko, Jonas Baltrusaitis

“Surface chemistry of carbon dioxide revisited”

Surface Science Reports, 2016, 71 (4), 595-671

Authors: William E. Taifan, Jean-Francois Boily, Jonas Baltrusaitis

“CH4 conversion to value added products: Potential, limitations and extensions of a

single step heterogeneous catalysis”

Applied Catalysis B: Environmental, 2016, 198, 525–547

Authors: William Taifan and Jonas Baltrusaitis

“Dairy wastewater for production of chelated biodegradable Zn micronutrient

fertilizers”

ACS Sustainable Chemistry & Engineering, 2016, 4 (3), 1722-1727

Authors: Hanyu Zhang, Megan Frey, Criztel Navizaga, Courtney Lenzo, Julian Taborda,

William Taifan, Abdolhamid Sadeghnejad, Alfredas Martynas Sviklas, Jonas Baltrusaitis

“Elucidation of ethanol to 1,3-butadiene reaction mechanism: Combined

experimental and DFT study”

Conference paper at Energy and Fuel – Biomass, ACS 2016 Fall Meeting at Philadelphia

Conference paper at Catalysis and Reaction Engineering Division, AIChE 2016 annual

meeting

Conference paper at North American Catalysis Society, 2017 annual meeting at Denver

Award and Honors

2016 Chevron Scholar Awards Lehigh Chemical Engineering

2017 Kokes Scholar Award – North American Catalysis Society

2017 John C. Chen Fellow – Lehigh University

Students Mentored (Past 4 Years)

• George X. Yan, B.S. Chemical Engineering, Lehigh University, Ph.D. Student, UCLA

• Paige Rockwell, B.S. Physics, Lycoming College

• Yiying Sheng, B.S. Chemical Engineering, Lehigh University